+ All Categories
Home > Documents > Pressure Loss Modeling of Non-Symmetric Gas Turbine ...

Pressure Loss Modeling of Non-Symmetric Gas Turbine ...

Date post: 11-Dec-2021
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
130
Pressure Loss Modeling of Non-Symmetric Gas Turbine Exhaust Ducts using CFD Steven Farber A Thesis in The Department of Mechanical and Industrial Engineering Presented in Partial Fulfillment of the Requirements for the Degree of Master of Applied Science (Mechanical Engineering)at Concordia University Montreal, Quebec, Canada April 15, 2008 © Steven Farber, 2008
Transcript

Pressure Loss Modeling of Non-Symmetric Gas Turbine

Exhaust Ducts using CFD

Steven Farber

A Thesis

in

The Department

of

Mechanical and Industrial Engineering

Presented in Partial Fulfillment of the Requirements for the Degree of Master of Applied Science (Mechanical Engineering)at

Concordia University Montreal, Quebec, Canada

April 15, 2008

© Steven Farber, 2008

1*1 Library and Archives Canada

Published Heritage Branch

395 Wellington Street Ottawa ON K1A0N4 Canada

Bibliotheque et Archives Canada

Direction du Patrimoine de I'edition

395, rue Wellington Ottawa ON K1A0N4 Canada

Your file Votre reference ISBN: 978-0-494-40910-7 Our file Notre reference ISBN: 978-0-494-40910-7

NOTICE: The author has granted a non­exclusive license allowing Library and Archives Canada to reproduce, publish, archive, preserve, conserve, communicate to the public by telecommunication or on the Internet, loan, distribute and sell theses worldwide, for commercial or non­commercial purposes, in microform, paper, electronic and/or any other formats.

AVIS: L'auteur a accorde une licence non exclusive permettant a la Bibliotheque et Archives Canada de reproduire, publier, archiver, sauvegarder, conserver, transmettre au public par telecommunication ou par Plntemet, prefer, distribuer et vendre des theses partout dans le monde, a des fins commerciales ou autres, sur support microforme, papier, electronique et/ou autres formats.

The author retains copyright ownership and moral rights in this thesis. Neither the thesis nor substantial extracts from it may be printed or otherwise reproduced without the author's permission.

L'auteur conserve la propriete du droit d'auteur et des droits moraux qui protege cette these. Ni la these ni des extraits substantiels de celle-ci ne doivent etre imprimes ou autrement reproduits sans son autorisation.

In compliance with the Canadian Privacy Act some supporting forms may have been removed from this thesis.

Conformement a la loi canadienne sur la protection de la vie privee, quelques formulaires secondaires ont ete enleves de cette these.

While these forms may be included in the document page count, their removal does not represent any loss of content from the thesis.

Canada

Bien que ces formulaires aient inclus dans la pagination, il n'y aura aucun contenu manquant.

ABSTRACT

Pressure Loss Modeling of Non-Symmetric Gas Turbine Exhaust Ducts using CFD

Steven Farber

In typical gas turbine applications, combustion gases that are discharged from the turbine

are exhausted into the atmosphere in a direction that is sometimes different from that of the

inlet. In such cases, the design of efficient exhaust ducts is a challenging task particularly

when the exhaust gases are also swirling. Designers are in need for a tool today that can

guide them in assessing qualitatively and quantitatively the different flow physics in these

exhaust ducts so as to produce efficient designs.

In this thesis, a parametric Computational Fluid Dynamics (CFD) based study was

carried out on non-symmetric gas turbine exhaust ducts where the effects of geometry and

inlet aerodynamic conditions were examined. The results of the numerical analysis were

used to develop a total pressure loss model.

These exhaust ducts comprise an annular inlet, a flow splitter, an annular to rectangular

transition region, and an exhaust stub. The duct geometry, which is a three-dimensional

complex one, is approximated with a five-parameter model, which was coupled with a design

of experiment method to generate a relatively small number of exhaust ducts. The flow in

these ducts was simulated using CFD for different values of inlet swirl and aerodynamic

blockage and the numerical results were reviewed so as to assess the effects of the geometric

and aerodynamic parameters on the total pressure loss in the exhaust duct. These flow

simulations were used as a data base to generate a total pressure loss model that designers

can use as a tool to build more efficient non-symmetric gas turbine exhaust ducts. The

resulting correlation has demonstrated satisfactory agreement with the CFD-based data.

m

Acknowledgments

I would like to first thank my supervisor Prof. W.S. Ghaly and Co-supervisor Ed Vlasic

for their guidance and suggestions on this project.

Thanks also go to the following people and organizations without which this work would

not have been possibe:

To Pratt & Whitney Canada for their financial and technical support, especially Remo

Marini and Mark Cunningham.

To the Natural Sciences and Engineering Research Council of Canada (NSERC) for

financial support.

IV

Contents

1 Introduction 1

1.1 Background 1

1.2 Single Port Annular-to-Rectangular Exhaust Ducts 3

1.3 Contribution and Scope of the Present Study 4

2 Theory and Literature Review 6

2.1 Diffuser Performance 6

2.1.1 Static Pressure Recovery Coefficient 6

2.1.2 Diffuser Effectiveness 7

2.1.3 Total Pressure Loss Coefficient 7

2.2 Conical Diffusers 7

2.2.1 Geometry 7

2.2.2 Swirl 8

2.2.3 Aerodynamic Blockage 9

2.3 Annular Diffusers 11

2.3.1 Swirl 14

2.3.2 Aerodynamic Blockage 19

2.4 Past Research Contributions 21

2.4.1 Loka et al 21

2.4.2 Cunningham 24

3 Design of Experiment 30

3.1 Geometric Design Space 30

v

3.1.1 Equivalent Cone Diffusion Angle 31

3.1.2 Flow Splitter Wedge Angle 31

3.1.3 Gas Path Aspect Ratio 34

3.1.4 Annular to Rectangular Transition Region 35

3.1.5 Exhaust Stubs 36

3.2 Aerodynamic Design Space 38

3.2.1 Swirl 38

3.2.2 Inlet Boundary Layer Blockage 40

3.3 Full Factorial Design 40

3.4 Taguchi Design 43

3.4.1 Assumption 43

3.4.2 Interactions 43

3.4.3 Selecting an Orthogonal Array 44

4 Geometry Synthesis 47

4.1 Equivalent Cone Diffusion Angle 49

4.2 Gas Path 51

4.3 Gas Path Aspect Ratio 51

4.4 Flow Splitter Leading Edge 51

4.5 Flow Splitter Wedge Angle 54

4.6 Duct Exit Cross-Section 54

4.7 Annular to Rectangular Transition Region 57

4.8 Plenum 59

5 Computational Study 62

5.1 Data Reduction 62

5.2 Pressure-Velocity Coupling 63

5.3 Advection Scheme 63

5.4 Turbulence Modelling 64

5.4.1 k - e 64

5.4.2 SST 64

VI

5.5 Computational Domain 65

5.5.1 Boundary Conditions 66

5.5.2 Grid Structure 68

5.5.3 Grid Study 72

6 CFD-Based Parametric Study 78

6.1 Effect of Swirl 79

6.2 Effect of Stub Direction 85

6.3 Effect of ECDA 87

6.4 Effect of Wedge Angle 89

6.5 Effect of Aspect Ratio and Area Ratio in the Annular to Rectangular Tran­

sition Region 92

6.6 Effect of Inlet Boundary Layer Blockage 96

7 Correlation of the Total to Total Pressure Loss 99

7.1 Japikse Correlation of Annular Diffusers 100

7.2 Correlation of the CFD Results 104

8 Conclusion and Recommendation 111

8.1 Conclusion I l l

8.2 Recommendation 112

Bibliography 114

vn

List of Figures

1-1 Pratt and Whitney PW200 engine 2

1-2 Pratt and Whitney PT6 engine 2

1-3 Exhaust duct geometric features 3

2-1 Conical diffuser presented with dimensional parameters 8

2-2 Conical diffuser performance chart based on data from Cockrell and Mark-

land (Bt « .20) 9

2-3 Conical diffuser performance chart from McDonald and Fox 10

2-4 Diffuser performance coefficient as a function of area ratio for (a) axial inlet

flow (b) swirling inlet flow 11

2-5 Radial distributions of total and static pressures at the inlet section . . . . 12

2-6 Variation of total pressure along stream surfaces of revolution 12

2-7 Maximum Pressure Recovery of Conical and Square Diffusers - Mth = 0.8 . 13

2-8 Variation of the total pressure loss coefficient with entrance length (X/D) for

a 5°; a) Reynolds number 1 x 105, b) Reynolds number 4 x 105 13

2-9 Conical diffuser loss map using Sharan's data 14

2-10 Exit discharge area ratio for conical diffusers on Cp* (based on data from

Cockrell and Markland) 15

2-11 Annular diffusers (a) equiangular (b) straight core (c) double divergent . . . 16

2-12 Annular diffuser performance chart (B\ ~ .02) 16

2-13 Pressure recovery contours 17

2-14 Straight annular diffuser performance with swirl 17

2-15 Performance of equiangular diffusers 18

viii

2-16 Loss map using Stevens and Williams data 19

2-17 Variation of the total pressure loss coefficient with entry blockage (Figure

adapted from Klein) . . 20

2-18 Exit discharge area ratio for annular diffusers on Cp* 20

2-19 Experimental setup 22

2-20 2D Cross-sectional View of Exhaust 22

2-21 Schematic of experimental setup 25

2-22 Domain of CFD grid used for computations 26

3-1 Equivalent Cone Diffusion Angle Levels Studied 32

3-2 Duct Wedge Angle Levels Studied 33

3-3 Gas Path Aspect Ratio 34

3-4 Gas Path Aspect Ratio Levels Studied 35

3-5 Duct Locations where Area is Specified 36

3-6 Exhaust Stub Direction 37

3-7 Sample of exhaust duct inlet swirl gradients 39

3-8 Boundary layer axial velocity profiles 41

4-1 Sample P&WC Single Port Swept Exhaust Duct 48

4-2 Equivalent Cone Diffusion Angle 49

4-3 Gas Path Conic 50

4-4 Gas Path Spline and Aspect Ratio 52

4-5 Flow Splitter 53

4-6 Flow Splitter Construction Plane 55

4-7 Flow Splitter Construction Plain 56

4-8 Flow Splitter Wedge Angle 57

4-9 Duct Exit Profile 58

4-10 Duct Cross-Sections Downstream of Flow Splitter 59

4-11 Duct Surface Passing Through Cross-Sections 60

4-12 Plenum 61

5-1 Air solid model of the computational domain 65

IX

5-2 CFD boundary conditions 67

5-3 Exhaust Duct Inlet Showing Prism Layer Elements 69

5-4 Plenum surface grid 70

5-5 Cross-section of the plenum domain showing element sizes 71

5-6 ^ p 1 ploted versus total number of prism layers 73

5-7 Three grids constructed from tetrahedral sizes of h/3 (coarse), h/6 (interme­

diate), and h/12 (fine) 74

APtt 91

5-9 Mach contours for the three grid densities . 76

5-8 ^P1 ploted versus total number of nodes 75

6-1 Cross-sections used for data reduction and flow visualization 78

6-2 Total pressure loss coefficient evaluated at each section defined in figure 6-1 79

6-3 Velocity contours normal to cross-section (nominal 0° swirl) 80

6-3 Velocity contours normal to cross-section (nominalO0 swirl)....con't 81

6-4 Velocity contours normal to cross-section (nominal 35° swirl) 82

6-4 Velocity contours normal to cross-section (nominal35° swirl)....con't . . . . 83

6-5 Mach contours at mid plane demonstrating the effect of stub direction (nom­

inal 0° inlet swirl) 85

6-6 Normalized static pressure contours at mid plane demonstrating the effect of

stub direction (nominal 0° inlet swirl) 86

6-7 Total pressure losses calculated for C-shape, Straight, and S-shaped exhaust

stubs (low inlet blockage) 87

6-8 Mach number contours at mid-plane demonstrating the effect of ECDA (inlet

conditions: nominal 0° swirl with low blockage) 88

6-9 Total pressure losses as a function of ECDA calculated at exit of the annulus 89

6-10 Wall shear stress contours (nominal 25° swirl) 90

6-11 Wall shear stress contours (nominal 35° swirl) 91

6-12 Total pressure losses as a function of wedge angle calculated at section 3 . . 92

6-13 Mach number contour plot on a plane through the flow splitter 93

6-14 Normalized wall static pressure contours demonstrating the effect of aspect

ratio and area ratio in the annular to rectangular transition region 94

x

6-15 Mach number contours demonstrating the effect of aspect ratio and area ratio

in the annular to rectangular transition region 95

6-16 Mach number contours demonstrating the effect of inlet blockage on three

exhaust ducts with ECDA of 0°, 10°, and 20° 97

6-17 Total pressure loss coefficient vs. inlet swirl for low and medium inlet blockage 98

6-18 Total pressure loss coefficient vs. inlet swirl for medium and high inlet blockage 98

7-1 Diffuser effectiveness versus area ratio with low level inlet aerodynamic block­

age and no inlet swirl 100

7-2 Diffuser effectiveness with principle geometric effects removed including data

at all levels of inlet aerodynamic blockage 101

7-3 Diffuser effectiveness with the principle effects of geometry and inlet swirl

removed according to preceding correlations 102

7-4 Diffuser effectiveness with the principle effects of geometry, inlet swirl, and

inlet blockage removed 103

7-5 Total pressure loss trend in the exhaust for low inlet blockage 104

7-6 Total pressure loss trend in the exhaust for medium inlet blockage 105

7-7 Total pressure loss trend in the exhaust for large inlet blockage 105

7-8 Surface passing through normalized losses for ECDA and aspect ratio (Solid

circles are points lying above the surface and hollow circles are points lying

below) 106

7-9 Normalized losses versus wedge angle with the effects of blockage, swirl,

ECDA, and aspect ratio removed 107

7-10 Normalized losses versus area ratio with the effects of blockage, swirl, ECDA,

aspect ratio, and wedge angle removed 108

7-11 Plot of CFD losses vs. predicted losses between inlet and section-12 . . . . 109

XI

List of Tables

2.1 Summary of performance of CFD analysis in the study of single port gas

turbine exhaust 28

3.1 Exhaust Duct Area Ratio (referenced to figure 3-5) 37

3.2 Geometric Parameters 42

3.3 Aerodynamic Parameters 42

3.4 Lg Orthogonal Array 45

3.5 Lg Orthogonal Array Expanded for Stub Direction 46

7.1 Break Down of Correlation Accuracy 108

xn

Nomenclature

Variables

A area (in2)

AB blockage area JA (l — |j) dA

AE effective area JA jjdA

AR area ratio

AS aspect ratio

Cp static pressure recovery coefficient P2Z_Pl

Cpi ideal pressure recovery coefficient 1 — - ^ j

D diameter (in)

Dh hydraulic diameter (in) = 4 — - £ % % £ £ "

E effective area ratio ^f-

g swirl parameter ° ;R/ _, ,

h annulus height (in) = r0 — r»

i? height (in)

A; turbulent kinetic energy

if total pressure loss coefficient t2Z- ll

L diffuser length (in)

M Mach number

P static pressure (psi)

Pt total pressure (psi)

q dynamic pressure (psi)

r radius (in)

T static temperature (R)

Tt total temperature (R)

u axial velocity (ft/s)

ug tangential velocity (ft/s)

U maximum axial velocity in cross section (ft/s)

xin

Variables

APts

Pa - Pa

Ptx - P*

Greek Symbols

a

5

e

V

9

P

UJ

average swirl angle

blockage ±f^ x 100

boundary layer thickness (in)

turbulence disipation rate

diffuser effectiveness -rr~ c, Pi

half cone angle

density (slug/in3)

turbulent frequency

Subscripts

1

2

duct inlet

duct outlet

inner

outer

Superscripts

mass averaged quantity

xiv

Chapter 1

Introduction

1.1 B ackground

Turboprop and Turboshaft gas turbine engines find their way in many aerospace and indus­

trial applications. Figures 1-1 and 1-2 give a schematic layout of such gas turbines where

each engine component is labelled. Careful review of these two figures demonstrates that

the engine layout results in two very different exhaust ducts. In the first figure, Fig 1-1,

an annular exhaust duct is found, where the flow and losses are rather well understood. In

the second figure, Fig. 1-2, the engine layout does not allow for an annular exhaust duct

resulting in the exhaust gases being redirected in a direction which differs from the inlet

direction. A quick overview of the open literature shows that there is very little knowledge

about these single port annular-to-rectangular exhaust ducts, be it flow physics, shape, de­

sign methods, performance, etc... Therefore, there is a need to study such exhaust ducts

which redirect the combustion gases in a direction that differs from the inlet direction.

Significant gains in gas turbine performance can be made by reducing the exhaust duct

loss. For example, consider a simple gas turbine cycle with a pressure ratio of 10 and at a

constant power output, a 1% drop in the absolute back pressure to the turbine will result

in a 1% improvement in the specific fuel consumption, therefore care should be taken to

design an exhaust duct that would minimize any pressure loss.

1

accessory compressor power gearbox compressor turbine turbine

reduction gearbox

combustion chamber

exhaust duct

Figure 1-1: Pratt and Whitney PW200 engine (source: www.pwc.ca)

exhaust duct

combustion chamber

centrifugal compressor

inlet screen

output shaft

power turbine compressor

turbine

axial compressor

accessory gearbox

Figure 1-2: Pratt and Whitney PT6 engine (source: www.pwc.ca)

2

Inlet Exit

Figure 1-3: Exhaust duct geometric features

1.2 Single Port Annular-to-Rectangular Exhaust Ducts

As was explained in the previous section, single port annular-to-rectangular exhaust ducts

find their use in gas turbine applications where the output shaft does not allow for an annular

exhaust duct application. In this situation, the exhaust gases are required to be redirected,

crossing over the output shaft, and diffused to ambient through a rectangular exhaust stub.

The distinct geometric characteristics of these exhaust ducts are an annular inlet followed by

a flow splitter, an annular to rectangular transition region, and a rectangular exhaust stub,

see Fig. 1-3. The annular diffuser function is to efficiently diffuse the flow to a low Mach

number before the gases enter the transitional region. Within the transitional region, the

gases are forced to make an aggressive 90° turn crossing over the power turbine shaft. It is in

this region that the lower inlet Mach numbers will result in a lower loss making it important

to obtain as much diffusion as possible in the upstream annular diffuser. A flow splitter is

located at the start of the transition region to provide guidance to the gases directing the

flow around the inner annulus surface. The exhaust gases enter a rectangular duct after

3

crossing the inner annulus surface where diffusion is continued and the exhaust gases are

directed to the ambient atmosphere. In aerospace applications, these exhaust ducts are

bound to be small in size, making efficient diffusion difficult and sometimes impossible.

One such application of this exhaust duct configuration is found in PT6 engines produced

by Pratt and Whitney Canada, Fig. 1-2. A PT6 engine has two spools, one for the gas

generator, and a mechanically independent power turbine output shaft. The exhaust duct

is mounted downstream of the last turbine stage. The flow into the exhaust duct is subsonic

and swirling with the magnitude of swirl varying over the engine operating range.

1.3 Contribution and Scope of the Present Study

In this work, a pressure loss model is produced for single port annular-to-rectangular exhaust

ducts. A parametric study was carried out using Computational Fluid Dynamics (CFD)

to simulate the flow numerically in exhaust ducts where key geometric and aerodynamic

parameters are varied and their effect on the total pressure loss is observed. Furthermore,

the duct geometry is approximated through a five parameter model, which was coupled

with a design of experiment method to generate a relatively small number of exhaust ducts

for numerical simulation. The resulting numerical data that was produced, has been used

as database to generate a total pressure loss model that designers can use as a tool to build

more efficient non-symmetric gas turbine exhaust ducts.

The scope of the present study has been divided into the following sections consisting

of:

1. Theory and Literature Review

• summary of past research on simple 2D conical and annular diffusers

• summary of past research on single port annular-to-rectangular exhaust ducts

2. Design of Experiment (DOE)

• n-dimension design space reduced to a five parameter model to approximate the

duct geometry

• definition of the key inlet aerodynamic parameters

4

• use methods of DOE to generate a relatively small number of exhaust ducts for

numerical simulation

3. Geometry Synthesis

• synthesis of a single port annular-to-rectangular exhaust duct based on the five

geometric parameters

4. Computational Study

• discussion on the use of a commercial CFD solver

• definition of the computational domain and grid structure

5. CFD-Based Parametric Study

• presentation and discussion of the results of the numerical study

6. Correlation of the Total to Total Pressure Loss

• review of a correlation produced by Japikse [7]

• correlation of the numerical data produced from the present work

7. Conclusion and Recommendation

5

Chapter 2

Theory and Literature Review

2.1 Diffuser Performance

2.1.1 Static Pressure Recovery Coefficient

The static pressure recovery coefficient is defined as the static pressure rise across the diffuser

divided by the inlet dynamic head:

Cp = 5^L (2.1)

For an incompressible and isentropic flow Bernoulli's equation can be used to define an ideal

pressure recovery coefficient in terms the Area Ratio:

C« = ! - ^ (2.2)

A useful expression for annular diffusers which relates the influence of inlet swirl a.\ to the

ideal pressure recovery is given by:

ri tan^ a\ + 1

6

2.1.2 Diffuser Effectiveness

A useful parameter for evaluating the performance of a diffuser is through the diffuser

effectiveness:

This relation relates the actual diffuser pressure recovery to the maximum potential pressure

recovery.

2.1.3 Total Pressure Loss Coefficient

The diffuser total pressure loss coefficient, which is the parameter focused on in this work,

is determined in the same manner as the static pressure recovery coefficient where the total

pressure difference between diffuser inlet and outlet is divided by the inlet dynamic head:

K^-J^ (2.5)

The total pressure loss coefficient for simple diffusers with uniform inlet and exit flow

conditions can be determined from Cp and Cpi through:

K = CPi - Cp (2.6)

2.2 Conical Diffusers

2.2.1 Geometry

Conical diffusers, Fig. 2-1 are commonly characterized by various dimensionless parameters.

Two of such dimensionless parameters are the dimensionless length, L/D\, and area ratio,

AR = A2/A1. The AR of a conical diffuser can further be described through the following

geometric relation:

A R = [ l + 2(L/£>i)tane]2 (2.7)

Various studies have been performed on conical diffusers that take account inlet condi­

tions as well as non-dimensional length and area ratio. One such study was performed by

7

Y

Figure 2-1: Conical diffuser presented with dimensional parameters

Cockrell and Markland and was presented as a performance chart, see Fig. 2-2, in a paper

written by Sovran and Klomp [1]. This chart presents pressure recovery coefficient (Cp)

versus non-dimensional length and area ratio. Furthermore the locus of the maximum pres­

sure recovery for both non-dimensional length (Cp*) and area ratio (Cp**) can be found.

Another study performed by McDonald and Fox [2] has produced a performance map with

similar results, Fig. 2-3.

2.2.2 Swirl

The performance of a conical diffuser can be affected significantly when a swirling component

is introduced to the flow field. The effects of swirl is different depending on the type of

swirl distribution (free vortex, forced vortex) effecting both the boundary layer and the core

of the flow. The importance of swirl on diffuser performance can be seen in the work of

McDonald et al. [9]. It is evident in Fig. 2-4 that the introduction of swirl can result in a

conical diffuser approaching the theoretical ideal performance.

Additional studies were performed by Senoo et al. [10], who studied conical diffuser

performance with swirling flow inlet conditions. The magnitude of swirl studied is quantified

1. I

.LLA

vC

_ — i

/ / rr

A *--/(*

*

£

_

Figure 2-2: Conical diffuser performance chart based on data from Cockrell and Markland (Bx « .20) [1]

through a swirl parameter defined as:

g = ^ ^ ^ ^ (2.8) R J0 (u2) rdr

These authors identified the development of a Rankin vortex type flow composed of a large

solid vortex core at the center of the conical diffuser where the axial velocity was very low.

The total and static pressure distribution for an inlet swirl parameter of g = 0.18 is shown

in Fig. 2-5. In Fig. 2-6, Senoo et al. have presented stream surfaces of revolution starting

from the axis to 100% at the wall. Total pressure is observed to rise in the core of the flow up

to a stream surface at 10% and again for 80% and 90%. This is due to the low momentum

flow at the core and the wall being dragged along by the main flow, and inversely, the low

momentum flow slows down the main flow producing a total pressure drop.

2.2.3 Aerodynamic Blockage

The effect of aerodynamic blockage in conical diffusers has been studied by many researchers.

One study by Livesey and Odukwe [13] looked at the length of the pipe preceding the conical

diffuser. Their results show that as the inlet pipe length is increased, the boundary layer

thickness increases resulting in a reduction in pressure recovery. A clear presentation of the

effect of aerodynamic blockage on pressure recovery was made by Dolan and Runstadler

9

Fi gure 2-3: Conical diffuser performance chart from McDonald and Fox [2]

[11] shown in Fig. 2-7. This study made a comparison between a conical diffuser and a

straight channel diffuser showing that the aerodynamic blockage is a significant aerodynamic

parameter. Sharan [12] carried out a study with careful measurements of pressure recovery

and total pressure loss on a 5° conical diffuser where he varied inlet pipe length, Reynolds

Number, and turbulence intensity. The results of his work, Fig. 2-8, demonstrate that

the total pressure losses increase with developing inlet blockage but following losses can

be seen to fall as a result the inlet flow field developing its own flow structure. Japikse

[14] has produced a loss map, see Fig. 2-9, using the data from Sharan where the data

loosely followed the trend of K = CPi - Cp demonstrating that inlet factors such as inlet

pipe length, Reynolds Number, and turbulence intensity must be considered. Kline [15] has

compared the results of numerous studies showing that approximately the same reduction

in pressure recovery for inlet aerodynamic blockage levels of approximately 14%. An early

correlation was produced by Sovran and Klomp [1] who have postulated that, for geometries

that are dominated by pressure forces (as oposed to viscous forces) the exit discharge can

be correlated with inlet blockage and area ratio. The results of their attempt to correlate

the data of Cockrell and Markland for conical diffuser geometries laying on the Cp* line is

presented in Fig. 2-10 demonstrating that the data collapse reasonable well.

10

-

" s

Symbcl

-o D

s

Divergence Angte. 29

4 .0 " 6 ,0 " i_

^ ^

a

,1 .1

-

-

-

6,03 8 - 0 « 1

0 12 .0 * , f M

0 15.8* ft 31.2"

Figure 2-4: Diffuser performance coefficient as a function of area ratio for (a) axial inlet flow (b) swirling inlet flow [9].

2.3 Annular Diffusers

Annular diffusers are geometrically represented similar to conical diffusers; however more

independent variables are present which require more complex relations. As in conical

diffusers, non-dimensional length is represented as, L/h, and area ratio, AR = A2/A1. The

AR of a three types of annular diffusers, shown in Fig. 2-11, can further be described

through the following geometric relation where:

1. Equiangular case

AR = l + 2(L//i)sin0 (2.9)

2. Straight Core is

AR = 1 + 2L sin 9 L2 sin2 0 / 1 - r ; / r0

h{\ + n/r0) + h2 1 + ri/r0

(2.10)

3. Double divergent

AR = 1+2 L\ s in0 1 + (rj/ro)sinG2 .(L\2 [l - n/ro] (sin2 ©i - sin2 0 2

h 1 + n/r0 -+lk 1 + n/ro (2.11)

11

100

40 0

K m,-0-I8

«-•? • • • *L

j * : Direct metnoii / ° : Present Tfth*)

K

0 10 40 60 16-75 r mm

Figure 2-5: Radial distributions of total and static pressures at the inlet section [10]

• Static fa^ssur* al.w$ th# cealfr • Stat* fswsAMt? sliwa ibe waU

3 ? ? / D ,

-A—A / / - • ,

| l l | < ' / T """—-Stal* irewc ilsnj ft* «n(er" — State pmswt ai»r.} the *8ll

(a )

2

0>}

3 z/D,

Figure 2-6: Variation of total pressure along stream surfaces of revolution [10]

12

*

8.3

Figure 2-7: Maximum Pressure Recovery of Conical and Square Diffusers - Mth = 0.8 [11]

0,4

0.3

o.2r. /

0.1

**\

o AR* 1.5 «4S = 2,5 „ 4ft* 3.5 n/W»S.O

0

SA-'-v<V * " ' tx_„.

^ ^•Y /

0 20 40 60 80 100 120 140 160

mo (a)

0.4 s < j 0.3[•

0.2!

0.1

0

o API- 1.5

<, 4fl«3.S n WJ«5.0

=:qr:i=^—-^

0 20 40 60 80 Too' 120 140 160 XID

(b)

Figure 2-8: Variation of the total pressure loss coefficient with entrance length (X/D) for a 5°; a) Reynolds number 1 x 105, b) Reynolds number 4 x 105 [12]

13

1

\

K° m 0

— ~ r —

*

» X

Symbol 0

E) f

m

r i -AR X/O Inlet 2.5

a.s

S.O

5.0

\

10-150

10-150

TO-150

10-150

—J

i Ha

1X10S

snias •

w o 5

6x10 B "

-

0.5 0.6 0,7 0.8 0.3 1.0

Figure 2-9: Conical diffuser loss map using Sharan's data [12]

Curved wall diffusers, which include axial to radial diffusers, are more complicated to

describe requiring their own derivation of L/h and AR specific to each shape.

Some of the first used annular diffuser maps produced by Sovran and Klomp [1] and

Howard et al. [3] were published in 1967, Figs. 2-12 and 2-13. Their research examined an

extensive selection of geometric diffuser types and produced detailed analysis of performance

measurements. The diffuser map presented in Fig. 2-12 shows the bulk of configurations

which gave the best performance. Same as with conical diffusers, the locus of the maximum

pressure recovery for both non-dimensional length (Cp*) and area ratio (Cp**) can be found.

The main difference between these studies is that the research of Howard et al. covered

fully developed inlet flow conditions while Sovran and Klomp covered low inlet blockage of

approximately .02.

2.3.1 Swirl

The effect of inlet swirl on pressure recovery has been studied by researchers and summarized

by Japikse and Baines [4] in Fig. 2-14. For each of the diffusers tested, a common trend has

been present. When inlet swirl is introduced the pressure recovery increases to a maximum

in range of 10° to 20° inlet swirl, and then pressure recovery decreases thereafter. The

effect of swirl on pressure recovery comes from two effects. The first is to press the flow

against the outer annulus surface due to the centrifugal force delaying separation on this

14

Bi-ora-- - 5**^-^

fis ** 2OQCO0

I^^C

J Cakoiati-cn

j

1

^o.?s

2 A 6 6 XJ

Figure 2-10: Exit discharge area ratio for conical diffusers on Cp* (based on data from

Cockrell and Markland)[l]

surface. The second comes from the centrifugal force destabilizing the inner hub boundary

layer resulting in the boundary layer approaching flow separation at the hub surface. From

the results shown in Fig. 2-14, it can be seen that the data of Coladipietro et al. shows that

equiangular diffusers are more efficient than the others. Elkersh et al. [5] studied equiangular

diffusers and confirmed the two effects that are produced are due to centrifugal forces acting

on the outer and inner annulus boundary layers. Their results show improvement in pressure

recovery up to inlet swirl values of 30°, and then decreasing performance with larger inlet

swirl values, Fig. 2-15. It is also demonstrated in Fig. 2-15 that the total pressure losses

tend to increase with increasing inlet swirl. A similar study to Elkersh et al. was performed

by Dovzhik and Kartavenko [16] on equiangular diffusers confirming that the total pressure

losses increase with increasing inlet swirl due to the intensity of flow separation at the outlet

along the inner hub. Klomp [17] has tested eight annular diffuser families where the inner

wall angles tested were both positive and negative. The results of this study demonstrated

that all diffusers tested were relatively insensitive to free-vortex type swirl ranging from 0°

to 25°. Greater inlet swirl levels lead to hub separation which was not found to result in

decreased performance in all diffusers tested. Swirl was found to have the largest impact

on the diffuser families with negative inner wall angles.

15

Figure 2-11: Annular diffusers (a) equiangular (b) straight core (c) double divergent

AR-1

as

\L^— £,,-02

0 4

'/ (J VA

• —

4^

1 2 5 10 20 Tfefl,

Figure 2-12: Annular diffuser performance chart {B\ ~ .02) [1]

16

Figure 2-13: Pressure recovery contours [3]

Sm."C&

0 ^ C A ' h ^ S1IM

<art$*o-,*n

n 'JVM'IX urn <d!'<m@&i<u

• J » t s t

4 i iSraaian sn3 Sr^sEaXa

X "6mlev ana *»j'4hes

_ i W a t ^ w M 61 ?

^ (Mwp-asro *>i 3.' •7 ujsi^p-eJra *fi,i ^ C* j {4<Jps«( t . ' ^< i '

^ . Gf~!3"A

>;

' , i

s

0

ft

t!

n /: >:• is

'<;

»iKJ5,

» £

§

* n

.>r! » .-.! SI

3 1

i. V

S41

8 5 .

" 1

•a

<<?«

•2 IB • J w

- :> ;f>

M

Aft

*>»SF

i I t

2 I"1

6 ?

4 =--7

, ,1

S i •5 t t j ,

' . • • 5

4JS

• ' t . ^ ,

l i cs

'• «6

{'•7

0 3 *

154

B S i

0»3

o«a ne? 0 7^

8

"*5ft

Hl£>>

S O D i

Uw*

t f j *

L, i . C M

SIC

E l f

t-T

r.

" t

» !,

e&WiiT%&i\im Maert ruirsji* ^ v~ s^>

Figure 2-14: Straight annular diffuser performance with swirl [4]

17

§ °3

o '•J>

•Si

© "Z 0.2

| -

"I 0-6

0.5

0,4

16° 20°

D

AR 1.65

'2.16 2.16

Predicted {!>

m 2:0 30 40 50

Swirl Made angle - j—i— " degrees

Figure 2-15: Performance of equiangular diffusers [5]

18

0 . 4 f

0.3

K 0.2

0 ,1

Q ^ _ __ __ ___ _^

Cp

Figure 2-16: Loss map using Stevens and Williams data [4]

2.3.2 Aerodynamic Blockage

The effect of thick ({3 — 0.10) and thin (/3 = 0.06) inlet boundary layer blockage with swirl

on annular diffuser performance was presented by Coladipietro et al. [18]. The authors

comment on the discovery of a forced vortex for the condition of a thick inlet boundary layer

partly due to the fact that the boundary layer penetrates deeply into the flow and meet near

the center of the annulus [18]. For the condition of a thin boundary layer, a forced vortex

is present near the wall but a free vortex is the predominant motion [18]. The authors

discovered that Cp was higher in diffusers with small non-dimensional length with thin

boundary layers and large non-dimensional length with thick boundary layers [18]. Japikse

[4] presented numerous data published by Stevens and Williams [19] in a study showing

the effect on inlet blockage, Fig. 2-16. From the data in this Fig. Japikse comments that

increasing inlet blockage results in reducing diffuser pressure recovery, however, when long

inlet lengths are present thus producing fully developed flow, the result is increasing pressure

~ N.

\ -.1 '\

\ — x \

V - Cp,

V

\ v-.a \

\*b*o \*9

\

„ l

Of0 a o , o 189 9x

0.101 \ J*

3

gNoraal I n l e t Boundary L»v«r Bloctcaq* Slew tttr&ilience)

• i n l e t Prid Used to fl«»«r#t« Higher I n l e t

Not*? ^ w ^F

28 \

o*ioV J*°-°«

\ \ !

S S

Turbulence Lab*lad nwnbers ~ indie*e« i n l e t feloekaew

**

1 X

19

W 2

0.10

OJ08

W*

&f:

f l / h , = SJ9J

Spoiler ^

Ni j

h I

W.I*

fi*9

fifW

w%

1 *

a-*"

i

ii/MUf

i** ^ **" v a

\

> ooa aioi cos oba aio o.i

Bf

Figure 2-17: Variation of the total pressure loss coefficient with entry blockage (figure adapted from Klein [6])

8 10

Figure 2-18: Exit discharge area ratio for annular diffusers on Cp* [1]

20

recovery. Klein [6] has taken the same data from Stevens and Williams and plotted it versus

inlet blockage, Fig. 2-17, showing dramatic improvement in the total pressure losses with

increasing inlet lengths. In the same manner as conical diffusers, Sovran and Klomp [1]

have added their test data on annular diffusers to the correlation presented in Fig. 2-10 for

conical diffusers, see Fig. 2-18, and found that for the inlet blockage tested the data are in

agreement for both geometric types.

2.4 Past Research Contributions

2.4.1 Loka et al.

A numerical and experimental study was carried out at Pratt and Whitney Canada to

achieve optimum integration of the PT6C-67A gas turbine engine on the Bell 609 aircraft

[20]. The authors conducted the numerical analysis using an in-house finite element, com­

pressible, Navier-Stokes CFD solver with &k — ui turbulence model. The efforts consisted of

three phases; the first was to optimize the uninstalled engine; secondly the numerical sim­

ulation was expanded to include the installation effects which included the exhaust ejector

system; lastly, experimental tests were conducted to validate the analysis.

Experimental Study

The experimental study was carried out on a full scale exhaust duct, Fig. 2-19. The

exhaust duct was mounted to a blower which could not attain the normalized flow levels

of an operating engine, therefore the authors had to extrapolate the data to represent

exhaust performance for an engine in flight. Inlet conditions were produced through a swirl

generator, which comprised of a series of adjustable vanes capable of producing swirl angles

in the range of 0° to 40°.

Computational Study

The computation domain consisted of a swept exhaust duct, an exhaust stub, and a plenum

chamber, Fig. 2-20. The plenum chamber was created to capture the sudden expansion

of the exhaust gases into the atmosphere. The computational boundary conditions at the

21

"SiCtis-;" •.

Figure 2-19: Experimental setup [20]

Farfield

FAfield Met, m as s flo w imp ose d (Air.p aft Flight) Farfield

Ps impo

v

Exit, ded

""""— 1 Exhaust Inlet, Enane mass flow imposed C enterline Engine Fl ow Pr cfil e

2;ure 2-20: 2D Cross-sectional View of Exhaust [20]

22

exhaust duct inlet were representative of engine profiles produced by the last stage turbine

rotor blades. Far-field boundary conditions in the plenum were modeled corresponding to

an aircraft in flight where, mass flow was imposed at the plenum inlet, and ambient static

pressure at the plenum exit.

Loss Mechanisms

Based on the CFD results, the authors [20] suggest that there are three pressure loss mech­

anisms:

1. Incidence on the flow splitter: The authors have observed a stagnation zone on the

suction surface of the flow splitter where the flow has separated. This stagnation zone

results in narrow layer of separated flow along the hub surface which merges with hub

wake.

2. A wake being shed from the hub surface due to the cross flow effects: The effect of the

flow crossing over the hub toward the exit port leads to creating a wake downstream

of the hub surface. The authors suggest the existence of a Von-Karmen vortex sheet

and evidence of two counter rotating vortices.

3. Excessive diffusion along the inner curve resulting in flow separation: A peak in Mach

number is found at the duct inner curve. The excessive diffusion in this region results

in the flow separating producing a pressure loss

From the three loss mechanisms, two can be identified at the duct exit plane by regions of

low total pressure; Hub separation and separation due to the inner curvature. It is suggested

that the size of the low pressure regions dictate the magnitude of each loss mechanism.

The authors have concluded that when there is no swirl at the turbine exit, the main loss

mechanisms are due to the hub separation and the inner curvature which have been assessed

to be equal contributors to the pressure losses. At higher swirl conditions, the incidence

along the flow splitter will lead to larger losses which can not be identified at the duct exit

because the separation merges with the hub wake.

23

Results and Conclusions

A comparison was made between the CFD and experimental results where the total-to-total

pressure loss coefficient, total-to-static pressure loss coefficient, and discharge coefficient are

compared. The authors conclude that the trends predicted by the CFD are the same as

what was found from rig testing, however the absolute levels varied between the two.

2.4.2 Cunningham

A detailed experimental and computational study was carried out on a single port tractor

exhaust duct at Queens University in cooperation with P&WC [8]. In this study, the objec­

tive was to determine the effect of inlet conditions and duct geometry on the flow structure

and the level of overall pressure losses. Conclusions were also made on the suitability of

boundary conditions for both experimental and computational work.

Experimental Study

The experimental study was carried out on a stereolithographic ^ scale model of the tractor

exhaust duct mounted to an annular cold flow wind tunnel, Fig. 2-21. A total of four

geometries were studied experimentally. The wind tunnel used was capable of producing

swirl, mass flow and inlet total pressure distributions similar to those seen in a gas turbine

engine. The range of swirl studied consisted of zero swirl and two radial profiles provided

by P&WC which are representative of what a sample PT6 engine exhaust duct would see at

the duct inlet plane. Total pressure profiling screens were used to produce circumferential

non-uniform total pressure profiles at inlet to the duct.

Computational Study

Five geometries were studied computationally. The computational domain consisted of an

inlet annulus, an exhaust duct, and a plenum chamber overlapping the exhaust duct exit,

Fig. 2-22. The plenum is a large conical domain with boundary conditions to allow the

exhaust jet to entrain flow freely into the plenum. The plenum also served to allow for

a non-uniform pressure distribution at the exhaust duct exit plane which results from the

large stream line curvature in the flow.

24

ffl n E ns

Figure 2-21: Schematic of experimental setup [8]

25

Plenum

7v \

*li\ . • • • • ' • . . . . ' . . . \

' • . . ' ; ; . . / . £'..••'„•-•' :'*.

• • ? * * . . " ' * ' • . * . - " . 4 . " *•

' * • . > ' * * J \

Inlet Annulus

• • - : : . . " • ' • • • i ^

••.•...-'•••-•• . " t r f c t i j

d u c t

Figure 2-22: Domain of CFD grid used for computations [8]

26

The computational grid was created using Gambit. Hexahedral elements were primarily

used to produce a structured mesh with only the annular to rectangular transitional region

requiring an unstructured grid composed of tetrahedral elements for ease of meshing. The

boundary layer was defined with prism elements in the first rows of mesh from the duct

surface. The flow solver used in this study was the commercial code Fluent 5.5. The most

suitable turbulence model which was available was the RNG k—e model; however the author

used the realizable k — e turbulence model through the majority of the study due to the

difficulty in obtaining a converged solution using the former.

Loss Mechanisms

Cunningham [8] has identified three geometric parameters affecting the total pressure losses

base on preliminary testing and literature:

1. Flow splitter.

2. Streamlining downstream of the center body.

3. Stub cross-sectional shape.

The three main pressure loss mechanisms were found and identified as:

1. Secondary flows: The secondary flows are generated through the duct bends as well

as the presence of the flow splitter redirecting the flow across the center-body.

2. Flow non-uniformity: Present at the stub exit representing undiffused kinetic energy

and therefore lower static pressure recovery. The exhaust stub cross-sectional shape

influenced the exit effective area-ratio.

3. Flow separation and recirculation: The total pressure losses are a function of the flow

separation and recirculation. Due to the complex shape of the exhaust duct, these

losses dominated over skin friction losses. Flow separation occurs along the inside

bend of the duct and in some cases, downstream of the center body.

27

Results and Conclusions

Cunningham [8] made a comparison of experimental and CFD results concluding that the

CFD consistently under-predicts the level of pressure losses in the exhaust ducts. CFD

has on the other hand shown that it is capable of capturing trends in losses by accurately

predicting changes in magnitude of pressure losses from one geometry to the next. When

comparing pressure losses due to inlet swirl, Cunningham found that the slope of the trend

line was under predicted when compared to measured results. This has been explained as

the inability of the turbulence models to handle the anisotropy of highly swirling flows.

Table 2.1 summarizes these findings.

Table 2.1: Summary of performance of CFD analysis in the study of single port gas turbine exhaust [8]

Parameter E

A P t t A Pts

geometry

flow structure

efficiency

inlet conditions

Suitability fair fair fair

good

fair

excellent

excellent

Comments Under-predicts distortion and secondary flow under-predicts losses under-predicts losses able to predict correct magnitude and trends with change in geometry easily gives details of internal flow structure, may not be reliable in identifying separation very efficient for studying inlet conditions, geometry limited by efficiency of mesher inlet conditions can be easily specified

If CFD is to be used to design optimum exhaust ducts, Cunningham has made the

following recommendations:

• In terms of predicting the total-to-static losses in the duct, the distribution of inlet

flow has a large effect on the distribution of the outlet flow. To be able to make a

realistic estimate of the total-to-static losses, a good estimate of the outlet flow is

required. To ensure this, velocity boundary conditions should be applied at the inlet

as this leads to a more realistic flow distribution at the exit.

28

• A large plenum is required at the exit of the duct to produce accurate flow distortions

in the duct near the exit.

• Turbulence models which account for swirl should be used if possible.

• Where possible, boundary conditions should be applied that account for the total

pressure non-uniformities at the duct inlet resulting from the presence of the engine.

29

Chapter 3

Design of Experiment

The work of Loka et al. [20], and Cunnigham [8] have identified many of the geometric and

aerodynamic parameters responsible for the overall duct loss. In this chapter, more param­

eters are identified. Also discussed here are how the geometric parameters are quantified

and bounded within a specific design space, giving limits to the magnitude of each param­

eter, for the purpose of creating a loss correlation. Next, combinations of each geometric

parameter are grouped together to create a set of exhaust duct models which, later, will be

simulated numerically along with the aerodynamic parameters to produce data for building

a loss correlation.

3.1 Geometric Design Space

A design space can be envisioned as being an n-dimensional box which is capable of contain­

ing all practical exhaust duct shapes and sizes. The size of the box is chosen to allow each

geometric parameter to be varied from a minimum to a maximum value which is thought

to cover rather well the design space so that both good and bad performing exhaust ducts

are represented. For each parameter, a minimum of three changes are required to be able to

predict a non-linear trend with respect to exhaust duct losses. In this study, each geometric

parameter will be extended to the minimum and maximum limits of the design space with

one selection in the center.

30

3.1.1 Equivalent Cone Diffusion Angle

The exhaust duct region between the inlet and the flow splitter can be represented as an

annular duct. Quantifying an annulus using one parameter is done through an Equivalent

Cone DifFuser Angle (ECDA). Equation 3.1 combines inlet annulus area, Ain, exit annulus

area, Aout and duct length, L, into one convenient parameter.

ECDA = 2 arctan M I Aou±

(3-1)

The range of ECDA chosen for the axisymmetric annular duct is to provide for attached

and separated flow. McDonald et al. [9] have found that optimal pressure recovery can be

found for conical diffusers at cone angles between 6° and 8° with decreasing performance

at larger cone angles. The work of Sovran and Klomp [1] has demonstrated that the

optimum cone angle is 8° provided that the area ratio is large between exit and inlet. It

was therefore decided that an ECDA of 20° should be sufficient to create flow separation

given that downstream of the annular duct is the flow splitter and annular to rectangular

transition region which could influence the streamwise pressure gradient and ensure that

flow separation will occur. It is also of interest to see the effect of no diffusion before

the annular to rectangular transition region where higher Mach numbers are expected to

produce higher pressure losses in this region, therefore the minimum ECDA studied is 0°.

An ECDA of 10° is used as a middle value as it was found to closely approximate the

sample P&WC duct and is close to the optimal cone angles found by McDonald et al [9].

Figure 3-1 shows the annular ducts which were used in this study upstream of the annular

to rectangular transition region.

3.1.2 Flow Splitter Wedge Angle

A swirling flow will create an incidence angle with the flow splitter leading edge possibly

leading to form a separated region along the suction surface. It is expected that small wedge

angles will lead to more severely separated flow resulting from larger incident angles. Using

the sample duct provided by P&WC as a the reference for a mid point value of around 45°,

min and max values of 10° and 80° were chosen for this study, Fig. 3-2.

31

.q..

20' 10* 0'

Figure 3-1: Equivalent Cone Diffusion Angle Levels Studied

32

10° Wedge Angle 45° Wedge Angle

80 Wedge Angle

Figure 3-2: Duct Wedge Angle Levels Studied

33

Gas Path Spline

-4 Figure 3-3: Gas Path Aspect Ratio

3.1.3 Gas Path Aspect Ratio

The annular to rectangular region of the exhaust duct includes a 90° bend where the flow

will have a strong tendency to separate from the inner curve. While it is not possible

to avoid making the 90° bend, it is possible to adjust the length of duct over which the

transition will occur. It is expected that a short duct will lead to severe flow separation

along the inner curve as a result of an increased pressure gradient. Longer ducts will have

lower pressure gradients reducing or delaying flow separation. The length of the annular to

rectangular transition region can be quantified through Eq. 3.2 which adds the duct height

as an additional parameter. This aspect ratio is the ratio of axial length, L, over radial

height, H, of a spline curve representing the general flow direction. The length and height

are measured between the start and end points of the gas path spline as demonstrated in

Fig. 3-3.

34

L/H — 0.92 l_/l — -| ^5

L/H = 1.38

Figure 3-4: Gas Path Aspect Ratio Levels Studied

AS = | (3.2)

For this study, the height has been fixed while the length varied. The mean gas path

aspect ratio used is 1.15 which is representative of the sample P&WC duct. To ensure

that flow separation occurs, an aspect ratio of 0.92 was selected as the minimum because

of the aggressive turn along the inner curve. An aspect ratio of 1.38 was selected as the

maximum with the expectation that flow will remain attached along the inner curve. Figure

3-4 demonstrates these aspect ratios where each cross-section was created with an ECDA

of 10° and the same exit area.

3.1.4 Annular to Rectangular Transition Region

An additional factor affecting the pressure gradient along the annular to rectangular tran­

sition is the remaining area ratio between the flow splitter and the duct exit, see Fig. 3-5

35

2 1

Figure 3-5: Duct Locations where Area is Specified

section 2 to 3. For a constant duct aspect ratio, any change in the area ratio will affect the

streamwise pressure gradient. Small area ratio ducts will have less diffusion and therefore

the smallest streamwise pressure gradient. Large area ratio ducts will have increased dif­

fusion, a large streamwise pressure gradient, and increase the possibility of flow separation

due to the increased boundary layer growth. The area ratios of the annular to rectangular

transition region have been selected to give, once combined with the annular region, Fig.

3-5 section 1 to 2, overall area ratios that do not exceed more then 2 from duct inlet to duct

exit, Fig. 3-5 section 1 to 3. Table 3.1 lists the combined area ratios of the annular section

with the annular to rectangular section to give overall duct area ratios.

3.1.5 Exhaust Stubs

Swept exhaust ducts can be found in turboprop and turboshaft configurations that require

the engine to be mounted to the aircraft with the output shaft pointed in the fore or aft

direction. The end result is to have the exhaust gases leave the engine through an exhaust

36

Table 3.1: Exhaust Duct Area Ratio (referenced to figure 3-5)

ECDA ARi_2 AR2_3 ARi_3

0° 10° 20° 0° 10° 20° 0° 10° 20°

1.000 1.000 1.000 1.180 1.180 1.180 1.450 1.450 1.450

1.000 1.175 1.350 1.000 1.175 1.350 1.000 1.175 1.350

1.000 1.175 1.350 1.180 1.387 1.593 1.450 1.704 1.958

Pusher Intermediate Tractor Figure 3-6: Exhaust Stub Direction

37

stub in a direction which will be dictated by the airframe manufacturer. Three exhaust stub

configurations are modelled in this study to account for the possible change in the duct losses

with stub direction. Two of the configurations turn the flow to create a Pusher (S-Shape

Duct) and Tractor (C-Shape Duct) application while the third takes the exhaust gases

straight out (Intermediate-Shaped Duct) defining a transitional point, Fig. 3-6. For each

exhaust stub configuration, the cross-sectional area and length was maintained constant,

while for the turning exhaust stubs, the turning radius, was maintained constant.

3.2 Aerodynamic Design Space

3.2.1 Swirl

Moderate inlet swirl has been shown to be beneficial in achieving optimum diffusion in two

dimensional diffusers. Annular diffusers can be designed with a large ECDA (ie. short ducts

with large diffusion angles), resisting flow separation when swirl is present at the diffuser

inlet. For the current study, the effect of swirl on diffuser performance is not as apparent as

it is in two dimensional diffusers. While inlet swirl should still be beneficial in the annular

portion of the exhaust duct, the flow splitter performance, however, will not benefit from

large incidence angles and the annular to rectangular region of the exhaust duct will result

in an asymmetric flow field.

Some inlet swirl angle profiles of several similar exhaust ducts can be seen in Fig. 3-7.

These profiles are produced at P&WC using an in-house code which neglects the upstream

effect of the downstream diffuser and predicts a circumferentially uniform flow distribution.

While many profiles can be observed in Fig. 3-7, it was decided that the swirl gradient used

in the current study would be of a constant gradient which fits within the given sample,

identified in Fig. 3-7 as the dotted line. This swirl gradient varies by 8° from r, to r0 and

is identified by the swirl value found crossing mid way along the annulus height, Fig. 3-7

shows nominal 0°. The range of nominal swirl angles selected to be studied varies from

nominal 0° to 35°.

38

Inlet Swirl Angle at Design Point

-2 0 2 Swirl Angle (°)

Figure 3-7: Sample of exhaust duct inlet swirl gradients

39

3.2.2 Inlet Boundary Layer Blockage

It is fully evident from Chapter 2 that inlet boundary layer blockage defined by

f — dA Blockage = A

fV,. x 100 (3.3)

JAdA

has a significant effect on diffuser performance. With regards to the diffusers in this study,

it is expected that the duct pressure losses will increase with increasing inlet boundary layer

blockage. The presence of an adverse pressure gradient will result in further increasing the

boundary layer thickness as flow travels downstream from the exhaust duct inlet. As the

boundary layer thickness grows, the wall shear stress reduces due to a decreasing velocity

gradient in the direction normal to the wall. Flow separation will occur as the wall shear

stress approaches zero, contributing further to the exhaust duct losses. A turbulent inlet

boundary layer will either prevent or delay the onset of flow separation. In the current

work, three values of boundary layer blockage have been studied and are displayed in Fig.

3-8. The largest inlet blockage expected to exist in practice is 10%. The presence of the

turbine upstream of the exhaust duct inlet will likely produce a turbulent boundary layer,

therefore the minimum expected inlet blockage will be less then 1%. An inlet boundary layer

blockage between 4% and 8% will give a good mean value for correlating inlet boundary

layer blockage later on in this study. The 1/7*'1 power law given by

has been used to define the velocity gradient within the boundary layer.

3.3 Full Factorial Design

To find the sensitivity of each of the geometric and aerodynamic parameters on exhaust duct

performance requires the modelling of numerous exhaust duct cases covering every possible

combination of parameters. This approach to performing a sensitivity analysis is termed a

Full Factorial Design requiring tremendous resources to complete a timely analysis. To put

this into perspective, this study has identified five geometric parameters having a first order

40

Inlet Boundary Layer 0,4

0.35 H

0.3 H

0.25 • Annulus Height

fr-rt) Jf^ 0 .2-

0.15 4

o.H

0.05 A

0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

u U

Figure 3-8: Boundary layer axial velocity profiles

41

Table 3.2: Geometric Parameters

Geometric Parameter Levels Number of Levels 1) ECDA 0°, 10°, 20° 2) Flow Splitter Wedge Angle 10°, 45°, 80° 3) AS 0.92, 1.15, 1.38 4) AR2_3 1.000 , 1.175, 1.350 5) Stub Direction S-Shape, C-Shape, Straight

Total Combinations =

3 3 3 3 3

243

Table 3.3: Aerodynamic Parameters

Aerodynamic Parameter Levels Number of Levels 1) Inlet Swirl Angle 0°, 10°, 25°, 35° 2) Inlet Boundary Layer Blockage Low, Med, High

Total Combinations —

4 3 12

effect on exhaust duct performance, where a minimum of three values are needed for each

parameter in order to find a non-linear effect on exhaust duct performance. Additionally,

each exhaust duct would need to be modelled using each combination of inlet swirl angle

and inlet boundary layer blockage. A full-factorial design with five parameters each having

three values will require 35 = 243 exhaust ducts to analyze as seen in Table 3.2. Each of

these exhaust ducts would then be modelled using each of the aerodynamic parameters in

Table 3.3 giving 12 aerodynamic boundary conditions. The total combination of geometric

and aerodynamic parameters which need to be modelled is 2916, which is an unpractical

task to perform. Consequently methods of selecting the minimum number of experiments

required to give the full information about each factor exist and are called partial-fraction

designs [21].

42

3.4 Taguchi Design

Dr. Genichi Taguchi [21] has created a set of guidelines for performing partial-fraction

designs using a special set of arrays called orthogonal arrays. In an orthogonal array all

levels of all factors are represented an equal number of times, and the combinations of any

two factors are also represented an equal number of times. In this respect, an orthogonal

array can be viewed as being well balanced, providing equal pairing between independent

parameters therefore reducing the total combination of parameters needed to determine

their effect on the dependent variable.

3.4.1 A s s u m p t i o n

Prior to choosing the orthogonal array, we must understand the assumption that is being

made in the Taguchi Design. The assumption is that the factors that are selected are

independent of each other and can be separated. Any interactions are assumed to have a

higher order effect on the dependent parameter and are confounded within the main effects.

The objective of this study is to catch the first order effects on exhaust duct losses. While

some interactions are expected to exist, they are assumed to have a smaller influence on

the exhaust duct losses then do the independent effects.

3.4.2 I n t e r a c t i o n s

Interactions that can be expected are upstream parameters affecting the downstream pa­

rameters, such as flow separation in the annulus on the flow splitter and the annular to

rectangular transition region. It is not expected that a separated flow will interact with the

downstream parameters at the end-walls the same way as an attached flow does because of

changes in the boundary layer, however the stream-wise momentum is going to be an order

of magnitude larger and will continue to relate stronger to the first order duct losses. There

is one particular interaction that should not be neglected in this study, namely the inter­

action between the exhaust duct and the exhaust stub direction because the same exhaust

duct can be used in both pusher and tractor configurations, Fig. 3-6.

43

3.4.3 Selecting an Orthogonal Array

An orthogonal array is selected by first considering the independent parameters one through

four in table 3.4. Each of the four parameters consist of three levels, therefore the minimum

number of geometries needed can be defined to be:

NV

NTaguchi = 1 + J2(L* ~ X) (3-5) i= l

Where:

NV = The number of parameters

and

Li = The number of levels for parameter i

For this study NV is equal to four and Li is equal to three for each of the parameters

giving nine geometries. An Lg orthogonal array shown in Table 3.4 is well suited for this

study. To include the interaction with exhaust stub direction, the L% orthogonal array

is expanded to allow each of the nine geometries to be combined with all stub directions

yielding 27 geometries to study. Table 3.5 presents the expanded Lg orthogonal array with

the physical variable to give exhaust duct families A through G. It can be well observed

that the Taguchi design has reduced the amount of geometries required for the sensitivity

analysis from 243 in the full factorial design to 27 in the partial factorial design.

44

Table 3.4: L9 Orthogonal Array

Independent Parameter Build Parameter 1 Parameter 2 Parameter 3 Parameter 4

1 2 3 4 5 6 6 8 9

1 1 1 2 2 2 3 3 3

1 2 3 1 2 3 1 2 3

1 2 3 2 3 1 3 1 2

1 2 3 3 1 2 2 3 1

45

Table 3.5: Lg Orthogonal Array Expanded for Stub Direction

Flow Sp: Duct ECDA Wedge I

~KA O5 io° A-2 0° 10° A-3 0° 10°^

T H O5 45° B-2 0° 45° B-3 0° 45°_

"CM 05 W C-2 0° 80° C-3 0° 80°_ D-l 10° 10° D-2 10° 10° D-3 10° HT E-l 10° 45° E-2 10° 45° E-3 10° 45°_ F- l 10° 80° F-2 10° 80° F-3 10° 80°

~~GA W 10° G-2 20° 10° G-3 20° 1(F H-l 20° 45° H-2 20° 45° H-3 20° 45°_

~ f l 20° 80° 1-2 20° 80° 1-3 20° 80°

ie AR2-3 AS Stub Direction

1.000 092 C-Shape 1.000 0.92 S-Shape 1.000 0.92 Straight 1.175 1.15 C-Shape 1.175 1.15 S-Shape 1.175 1.15 Straight 1.350 1.38 C-Shape 1.350 1.38 S-Shape 1.350 1.38 Straight 1.175 1.38 C-Shape 1.175 1.38 S-Shape 1.175 1.38 Straight 1.350 0.92 C-Shape 1.350 0.92 S-Shape 1.350 0.92 Straight 1.000 1.15 C-Shape 1.000 1.15 S-Shape 1.000 1.15 Straight 1.350 1.15 C-Shape 1.350 1.15 S-Shape 1.350 1.15 Straight 1.000 1.38 C-Shape 1.000 1.38 S-Shape 1.000 1.38 Straight 1.175 0.92 C-Shape 1.175 0.92 S-Shape 1.175 0.92 Straight

46

Chapter 4

Geometry Synthesis

The approach discussed here for geometry synthesis has been designed to produce exhaust

ducts which can be specified using only key geometric parameters thought to be responsible

for the total pressure losses in the exhaust duct. The 3D nature of the single port swept

exhaust duct makes its complete geometric representation impossible using only one dimen­

sional geometric parameters. Therefore assumptions are made to complete the geometry.

The blanks in the steps discussed in this Chapter have been filled in based on the assump­

tion of how a single port swept exhaust duct should be represented geometrically, and could

vary with each designer who uses this approach. It is the assumption of this author that the

steps not discussed here would only represent a second order effect on exhaust duct losses,

and therefore do not play a crucial role in the current study.

A sample of a single port swept exhaust duct taken from P&WC is shown in Fig. 4-1.

This exhaust duct has smooth flowing features which do not contain any sharp edges which

can disturb the flow. To describe the geometric features of this duct would require complex

splines and curves which do not suit the present study because of the numerous parameters

that would be needed. Steps have been taken to simplify the exhaust duct such that it can

be described by simple one dimensional parameters identified in Chapter 3.

47

Figure 4-1: Sample P&WC Single Port Swept Exhaust Duct

48

Figure 4-2: Equivalent Cone Diffusion Angle

4.1 Equivalent Cone Diffusion Angle

From examination of the annular region of the sample exhaust duct, the inner and outer

walls are not straight walled and are found to be represented by splines. The curved

profile has been made linear, and is now better suited to be described by Eq. 3.1. Further

examination of the sample exhaust duct shows tha t the hub surface is nearly straight walled

with a constant radius and has therefore been replicated this way in the current study. The

cross-sectional profile of the annular region is given in Fig. 4-2 which presents the sample

geometry with the simplified representation.

49

Figure 4-3: Gas Path Conic

50

4.2 Gas Pa th

The remaining portion of the gas path continues from the exit of the annular section to

the duct exit. The profile of the sample duct can be closely reproduced with the use of

conies as shown in Fig. 4-3. The conies used require five parameters; a start point, an

end point, a tangent direction at the start point, a tangent direction at the end point,

and a conic parameter. The start points and tangent directions are set from the preceding

section defined by the ECDA. The end points are defined by the exhaust duct exit area

and position. The tangent direction for the end points is always 90° from the duct inlet. A

conic parameter of 0.45 has been chosen based on a good fit with the sample duct and used

throughout this study. Once the inner and outer conies have been defined, equally spaced

points can be located along each curve and joined with straight lines. A gas path spline

can be constructed by locating the midpoints of each line and then connecting them with a

spline Fig. 4-4.

4.3 Gas Pa th Aspect Ratio

Once the inner and outer conies have been defined, equally spaced points can be located

along each curve and joined with straight lines. A gas path spline can be constructed by

locating the midpoints of each line and then connecting them with a spline, Fig. 4-4. The

gas path aspect ratio can now be measured according to Eq. 3.2.

4.4 Flow Splitter Leading Edge

The flow splitter is located downstream of the annular duct section at the bottom dead

center of the exhaust duct, Fig. 4-5). It can be described as being an axisymmetric vane

with an elliptical leading edge varying from hub to tip centered along a plane not fully

normal to the axial direction. Large fillets are used at hub and tip of the vane to merge

smoothly with rest of the domain. The flow splitter smoothly blends from the leading edge

outward to the exhaust duct through the annular to rectangular duct transition region.

To construct the flow splitter, a plane is located normal to the gas path spline slightly

51

Gas Path Spline

"ft Figure 4-4: Gas Path Spline and Aspect Ratio

52

Flow Splitter

Figure 4-5: Flow Splitter

53

downstream of the annular duct segment, Fig. 4-6. With the height of the flow splitter

located at the bottom dead center of the duct, it's center is located and a circle is drawn

of diameter D in a plain normal to the line defining the flow splitter height. The circle is

centered along the symmetry plane of the duct but not constrained to be centered on the

line defining the flow splitter height. Two lines are drawn at an angle a intersecting the

symmetry line of the flow splitter and tangent to the circle. The circle is allowed to move

along the center line with the tangent points falling on the plane positioned normal to the

gas path spline. The portion of the circle lying downstream of the plane positioned normal

to the gas path is removed leaving the upstream portion to be the leading edge of the flow

splitter, Fig. 4-7. The leading edge curve is then extruded to produce the flow splitter

which is then joined to the upstream annulus through fillets.

4.5 Flow Splitter Wedge Angle

The transition of the exhaust duct geometry from annular to rectangular continues from the

leading edge of the flow splitter through a wedge shaped inner passage directing the flow

around the hub toward the exhaust duct exit. The P&WC sample duct shown in Fig. 4-8

demonstrates that the varying leading edge diameter leads to a varying wedge angle from

hub angle a\ to shroud angle a^- The exhaust ducts created in this study contain a single

wedge angle as a result of using one leading edge profile from hub to shroud. To smoothly

merge the wedge angle into the duct transition section a limit was put on the flow splitter

length to allow a smooth transition to occur as seen in the P&WC sample exhaust duct in

Fig. 4-8. This limit was taken as being l/8 i fe the length, L, defined in the gas path aspect

ratio.

4.6 Duct Exit Cross-Section

The sample P&WC exhaust duct consists of an exit cross-section that is only symmetric

across one plane as shown in Fig. 4-9. The duct exit cross-section of the sample P&WC

exhaust has been developed through optimization for a given set of flow conditions which

are unknown to this author. For this reason, a fully symmetric exit cross-section, shown as

54

Gas Path Spline

Plane Normal To Gas Path Spline

Flow Splitter Height, L Center of Flow Splitter

Figure 4-6: Flow Splitter Construction Plane

55

Leading Edge

Tangents

Duct Symmetry Plane

Plane Normal to Gas Path Spline

Figure 4-7: Flow Splitter Construction Plane

56

P&WC Sample Duct Simplified Duct Wedge Angle

Figure 4-8: Flow Splitter Wedge Angle

the blue profile in Fig. 4-9, has been used throughout this study so not to introduce any

unknown influences produced by the sample P&WC exhaust duct.

4.7 Annular to Rectangular Transition Region

The shape of the exhaust duct from the flow splitter downstream to the exhaust duct

exit is defined by cross-section profiles built on planes passing through each of the lines

connecting the inner and outer conies shown in Figs. 4-3 and 4-4. The exhaust duct

transition from annular to rectangular produces complex cross-sections which cannot be

defined with straight lines and curves, resulting in the decision to use splines, Fig. 4-10.

Numerous control points are required to define each cross-section spline making it difficult

to develop a consistent approach to the design; however, some rules have been created and

followed throughout this study. The following rules that have been applied are:

1. The profile is bounded by the annulus hub radius.

2. The profile is bounded so as not to surpass the gas path inner conic.

57

Before

After

f ^

Symmetry Plane

(Before and After)

Symmetry Plane

(Only After)

Figure 4-9: Duct Exit Cross-Sectional Shape

58

Plane Downstream of Flow Splitter

Duct Exit

Figure 4-10: Duct Cross-Sections Downstream of Flow Splitter

3. The growth in cross-sectional area should be nearly linear.

4. The exhaust duct geometry should transition smoothly without any waviness or sharp

edges that may affect the flow of fluid.

To maintain the wedge angle defined in Sec. 2.2, a cross-section is placed passing through

a point marking the end of where the wedge angle is held to. Once the cross-sections are

completed, a surface is passed through them and then joined to the upstream flow splitter

and annulus, Fig. 4-11.

4.8 Plenum

A plenum domain, shown in Fig. 4-12, is created for each duct series as a function of the

exhaust s tub exit hydraulic diameter. The inlet surface parallel with the stub exit plane

has a diameter of 10Dhstubexit and the plenum length is \hDhstubexit- The half cone angle

of the plenum is 30°.

59

Figure 4-11: Duct Surface Passing Through Cross-Sections

60

30°

0 = 1OXDh stub exit

15 X D h s t u b e x i t

Figure 4-12: Plenum

61

Chapter 5

Computational Study

The computational analysis has been performed using CFX 5.7.1, which is a a commercial

Computational Fluids Dynamics (CFD) package developed by ANSYS. CFX 5.7.1, solves

the unsteady Navier-Stokes equations in their conservative form. The computational domain

is discretized into finite control volumes using a mesh. Each of the governing equations are

integrated over a control volume, such that the related quantity (mass, momentum, energy

etc.) is conserved in a discrete sense. The following sections cover the development of the

computational methodologies and mesh synthesis.

5.1 Data Reduction

Data reduction techniques for averaging flow properties must be carefully selected consid­

ering the non-uniformity of the flow through the exhaust ducts considered in this study.

Errors can be introduced through averaging techniques which will affect the performance

parameters introduced in Sec. 2.1, where the total pressure loss coefficient is of prime

interest in the current analysis.

Perhaps the simplest form of averaging flow parameters (</>) is the area average given

by:

Parea = ^ I 5 - 1 )

A more apropriate averaging method for non-uniform flows is the mass average method

62

given by: J((f>pudA)

* = 7 O ^ A T (5'2)

Wyatt [22] has studied the errors generated using averaging methods that include Eqs. 5.1

and 5.2 in ducts with various velocity profiles. It was demonstrated that total pressure

losses in a duct for a Mach number of 0.3 resulted in an error of less then . 1 % and 1%

for the mass average and area average techniques respectively. Based on the above results

and to be consistent with the work of Cunningham [8], all flow parameters have been mass

averaged in the present study.

5.2 Pressure-Velocity Coupling

To overcome the decoupling of pressure and velocity, CFX 5.7.1 uses a single cell, un-

staggered, collocated grid. The continuity equation is a second order central difference

approximation to the first order derivative in velocity. A fourth derivative in pressure is

used to modify the equation to redistribute the influence of pressure and overcome the prob­

lem of checker board oscillations found when variables are collocated. The method used is

similar to that from Rhie and Chow [23]. A number of extensions are implemented in CFX

which improve the robustness of the discretisation when pressure varies rapidly.

5.3 Advection Scheme

CFX 5.7.1 offers the first order upwind differencing scheme, the high resolution scheme,

or a specified blend factor to blend between first and second order advection schemes to

calculate the advection terms in the discrete finite volume equations. The high resolution

scheme, used in this study, has a blend factor which varies throughout the domain based

on the local flow field. In flow regions where there are low gradients, the blend factor will

take on a value close to one representing a second order advection scheme. Flow regions

with large gradients will have a blend factor near zero representing a first order advection

scheme to prevent overshoots and undershoots and maintain robustness.

63

5.4 Turbulence Modelling

A wide range of turbulence models are available in CFX 5.7.1 which include a standard

k-e, Large-Eddy-Simulation, and the Shear Stress Transport (SST) model. The two most

appropriate turbulence models for this analysis are the k-e and SST turbulence models.

5.4.1 k - e

The k-e model is robust and computationally inexpensive at solving turbulent flows. The

downfall of this turbulence model is that it predicts the onset of flow separation late and

tends to under-predict the magnitude of separation. Separated flow has been observed

in single port swept exhaust dust by Loka et al [20] and Cunningham [8]. Both authors

have observed flow separation along the exhaust duct hub and inner curve in the annular

to rectangular transitional region. It is important to note that diffusing flows produce

unstable boundary layers due to the flow traveling against an adverse pressure gradient.

The effect of predicting the onset of flow separation late results in over predicting exhaust

duct efficiencies, where flow separation leads to less pressure recovery and lower discharge

coefficients.

5.4.2 SST

A turbulence model developed to address the deficiencies of the k — e model is the SST

turbulence model developed by Menter [24]. This model works by solving the k—ui equations

at the near wall region and then the k — e in the free stream region with a blending function

for transition between the two models. The k — UJ based SST model takes into account the

transport of the turbulent shear stress to give highly accurate predictions of the onset and

the amount of flow separation under adverse pressure gradients. The key reason for the

deficiencies of the k — e model is that is does not account for the transport of the turbulent

shear stress resulting in an over prediction of the eddy-viscosity.

CFX guidelines for using this model requires an overall y+ of less then two and no less

then 10-15 grid points within the boundary layer. The SST model continues to be accurate

when these guidelines can not be achieved from the use of scalable wall functions. For fine

64

Figure 5-1: Air solid model of the computational domain

grids the k — to model is applied, but when a coarser grid is used, the model switches to a

wall function treatment making use of the logarithmic profile assumptions. For this reason,

a new near wall treatment was developed by CFX for the k — u> based models that allows

for a smooth shift from a low-Reynolds number form to a wall function formulation.

5.5 Computational Domain

The computational domains studied in this work are composed of an exhaust duct, exhaust

stub, and a plenum chamber as illustrated in Fig. 5-1. The plenum chamber has been

modeled to serve multiple functions which include providing a far-field boundary condition

and a domain for jet flow entrainment. The work of Loka et al, and Cunnigham have

made use of plenum chambers in their work, both demonstrating that non-uniform pressure

65

gradients exist at the exhaust stub exit. These pressure gradients are found to exist as

a result of the swirling flow and geometric influences causing streamline curvature. If no

plenum is modeled, the flow structure at the exhaust stub exit would be strongly influenced

by the imposed boundary conditions. A realistic exit flow structure can only then be

achieved if the pressure gradients at the exhaust stub exit are known in advance.

5.5.1 Boundary Condit ions

The CFD boundary conditions have been selected as being the most appropriate for deter­

mining the aerodynamic and geometric parameters sensitivity on exhaust duct losses. The

boundary conditions are not consistent with what is experienced at flight conditions where

ambient conditions change according to aircraft speed, altitude, and engine installation.

The following boundary conditions can be seen in Fig. 5-2.

Duct Inlet

Ansys CFX contains a large variety of inlet boundary conditions to suit many situations.

The requirements of this study are to impose a swirl gradient and aerodynamic blockage

at the duct inlet plane. Total pressure and mass flow boundary condition do allow for an

imposed flow direction, however, the aerodynamic blockage is an implicit result of the flow

simulation making these boundary conditions a poor choice for the current study. The

most suitable inlet boundary condition for this study, that is offered in CFX, is to impose a

velocity gradient with magnitude and direction. Using this option both a swirl gradient and

blockage can be controled explicitly. Using the later choice of inlet boundary conditions,

the velocity magnitude and direction were specified as a function only of radius (constant

across the circumference). An inlet total temperature of 1530 R was used throughout this

study, being a common average total temperature for P&WC swept exhaust ducts. The final

parameters set at the inlet to the exhaust duct are turbulence intensity and eddy viscosity

ratio. A turbulence intensity of 10% and an eddy viscosity ratio of 300 were suggested by

P&WC experts and used throughout the study.

66

Plenum Inlet:

Mass Flow = 10% iniet mass flux

Tt = 519R

1 = 1%

e = 1

A

Openinq:

Ts

Ps

l =

e =

= 519R

= 14.7 psi

1%

1

Inlet:

Velocity

•Magnitude

•Direction

Tt=1530R

I = 10%

£ = 300

Walls:

No Slip

Adiabatic

Figure 5-2: CFD boundary conditions

67

Plenum

The conical plenum domain has an inlet and an outlet boundary condition. The inlet

boundary condition is mass flow with direction. The mass flow rate is implicitly defined

to be 10% of the duct inlet mass flux. The flow direction over the full surface is directed

normal to the exhaust stub exit plane. A uniform total temperature is imposed at 519

R. A low inlet turbulence intensity and eddy viscosity ratio of 1% and 1 respectively was

used. The plenum exit was modeled as an opening allowing both inflow and outflow from

the same location. The opening static pressure and temperature is set at 14.7psi and 519

R respectively. For the condition of inflow through the opening, a low turbulence intensity

and eddy viscosity ratio was selected to match the plenum inlet boundary condition.

Walls

The exhaust duct and stub wall have been modeled as no slip and adiabatic.

5.5.2 Grid Structure

The computational grid was created using ANSYS ICEM CFD. This meshing package offers

the capability to create grids in multi-block structured, unstructured hexahedral, tetrahe-

dral, hybrid grids consisting of hexahedral, tetrahedral, pyramidal and prismatic cells. For

this study, the computational domain was constructed using unstructured tetrahedral ele­

ments with prismatic cells for near wall turbulence model requirements.

The three components of the computational domain (exhaust duct, stub, and plenum)

where meshed separately and later assembled in CFX making use of General Grid Interface

(GGI) which is used to create a fluid-to-fluid interface layer between two grid surfaces which

do not have matching node locations. The GGI theory involves a control surface treatment

of the numerical fluxes across the interface. The handling of the interface fluxes is fully

implicit and fully conservative in mass, momentum, and energy. The advantage of using

GGI capability of ANSYS CFX, is that a parametric study can be carried out with minimal

time spent creating computational grids. For one duct series in Chapter 3, only one exhaust

duct, one curved stub, one straight stub, and one plenum is needed to be meshed in ICEM

CFD. The C-Shaped domain is created by assembling the exhaust duct, curved stub, and

68

Figure 5-3: Exhaust Duct Inlet Showing Prism Layer Elements (20 layers shown)

plenum in ANSYS CFX and defining all the fluid-to-fluid interface surfaces. The S-Shaped

duct is created in ANSYS CFX by rotating the stub and plenum about an axis. To create

the Intermediate-Shaped duct, the curved stub is swapped with the straight stub, and the

plenum is rotated and translated in space to match interface surfaces.

Exhaust Duct and Stub Grid

The exhaust duct and stub grids are composed of prism elements to resolve the boundary

layer and tetrahedral elements to resolve the free-stream flow. These two computational

domains must respect the mesh requirements of the SST turbulence model discussed in

Sec. 5.4.2 since the walls in these domains are defined in CFX as no-slip surfaces where a

boundary layer is present. Prism elements are inflated from the walls of each domain up to

69

Figure 5-4: Plenum surface grid

a height which meets minimum requirement of putting 10-15 nodes in the boundary layer,

see Fig. 5-3. With the inlet area of the exhaust duct fixed in this study the prism layers

have been extended to a height of /i/6 measured normal from the wall surfaces. The prism

elements are formed by first growing one layer to the overall inflation layer height, xn, then

subdividing them using an exponential growth law with a height ratio of d, giving a smooth

transition of the last prism layer height, xn_i, to the tetrahedral elements resolving the free

stream flow. The height of a given prism layer, i, can be calculated from Eq. 5.3.

Xi = xix G P " 1 (5.3)

70

(jsissis^si e s is >s s» si s 5*3

& * * * * IS * * IS IS S IS a s !>QXlXIXIXlXlXlXlXIXIXlXIXP<f, e •:< s s s s si s s s s * s is s>s i$ e« si s g? m m s< si © m si si s e? is s< a s c © si s is is si si s is is si si s is is is s s S3i J S 3 © >S S •!< IS IS >S S >S S S S S S S 65 * S @ S S * S SfSe

@ si * is >s s >s s is K s >s £ <«* «•> >n s n< m n K< IS HOS IS (S IS IS S IS IS IS K£;?•

oxixixixixixixixixr ^ ^ S S ^ i S ^ ^ S ^ ^ ^ S ^ : ^ ; ^ . ^ ^ ^ . ^ : ^ ^ ^ : ^ : ^ : - : : - - ^ G l S S i S i S l S i S S i S ^ S ^ i ^ . * ^ - - ^ : ' ^ ©SS^Sf f lSSQff l iSOi©*; ; : :^ I e IS S S iS IS lS IS S IS lS IS 12^ :^ ;:*::: S;; ?* ::: :<*• ? £::: *::;: t::i::::^:-l&iS^S^diS^SlSlSiSl©&^: : - :^ :^-^:^: i ; -^- . ; -^ : : : - ; " -

^iSSIS^^^SIS^ISiSe^^iS.::.;:^^.^^:::'::-'::.^'ii1V^>i gi&oioioio^BiiojBBaK^:::;;.;:;.:::::;';',:;::;;-11

^ S d ^ S i S d d i S S d d S ^ ^ o ^ ^ ' l ^ ' i ^ ' V ^ ^ i i l S ^ S i 1 : SiSd^SiSddSiSdiS©^?J i:; :-^^'53.^fi! ie!dSlSlS^iS £< is >s si s is is si si s >s s s>* •; >-• a >n s s is is * s © * *

e f f iSS IS f f iBMSISSBSf f iSBBMWSlS lS lSEVV |& H K H S IS >S v IS S IS IS IS IS H IS IS IS IS S IS IS »3Q 6<iSSSSSiSiSiSSdiSKiSS>SiSiSSlSK?£^ e is is s s is ISIS <s s »is is si s is s s is vi

s is si s s >s is is sa s is si is si s Q g :i< m is x s is is is is s is is is & E S T

:;:;:;::;:; ';:v.: . iS & i S IS €

? ? ¥ h- S. *JSi IBs SI Si S IS S •!< g ^ v a Si S •!< IS SI SI S © ® * «

Figure 5-5: Cross-section of the plenum domain showing element sizes

71

Plenum Grid

The plenum domain is meshed with the inlet and opening surfaces as was explained in

Sec. 5.5.1. Only tetrahedral elements are used in this domain because there are no no-

slip surfaces requiring the grid resolution needed to resolve the boundary layer. One more

surface is included in this domain, shown in Fig. 5-4, having the same cross-sectional shape

as the exhaust stub exit plane. This surface is included for the GGI interface boundary

condition used to create the fluid-to-fluid interface which links the flow from the exhaust

stub to the plenum. The elements on this surface are sized to match the exhaust stub

tetrahedral size. The remaining surfaces do not require fine elements because high gradient

are not expected in these regions. The elements on these remaining surfaces are created

eight times larger then what is used in the exhaust duct and stub grid. Two ease the

transition from the small elements in the exhaust stub domains to the large elements in the

plenum, a volume of elements four times larger then the exhaust stub are produced in the

region where the jet flow first enters the plenum, Fig. 5-5.

5.5.3 Grid Study

Prior to fixing the mesh parameters for the current study, a sensitivity analysis has been

preformed to select the appropriate quantity of prism layers and tetrahedral grid density.

Selecting the Quantity of Prism Layers

Three sets of exhaust duct and stub meshes were created to test the sensitivity of the total

to total pressure losses versus the quantity of prism layers within an overall prism layer

height of h/6. For the three sets of meshes, total prism layer quantities of n = 20, 25,

and 30 were studied while the height, x\, of the first node from the wall was maintained

constant. Each of the meshes was created using the same density of tetrahedral elements

and same CFD boundary conditions so as to isolate the differences in numerical solution

with respect to the quantity of prism layers. The results of this analysis are plotted in Fig.

5-6. It is apparent that increasing the number of prism layers has only a small effect on the

total pressure losses within the exhaust duct and stub. Total pressure losses varied by only

2% when going from 20 to 30 prism layer. Considering the small differences in pressure

72

20 25

Total Prism Layers

Figure 5-6: ^-^ ploted versus of total number of prism layers

73

Coarse Intermediate Fine

Figure 5-7: Three grids constructed from tetrahedral sizes of h/3 (coarse), h/6 (intermedi­ate), and h/12 (fine)

losses, 20 prism layers have been used thought this study resulting in a 25% reduction in

nodes versus the 30 prism layer runs.

Selecting the Grid Density

Grid dependence on grid density has been examined by constructing computational domains

using different tetrahedral element sizes. Three domains were created using tetrahedral sizes

of h/3, h/6, and h/12 while maintaining the same prism parameters as selected in Sec. 5.5.3.

Figure 5-7 demonstrates that the selected tetrahedral parameters lead to a coarse mesh of

135465 nodes, an intermediate mesh of 322371 nodes, and a fine mesh of 980284 nodes. The

three grids were solved numerically using the same boundary condition used in Sec. 5.5.3 and

74

200000 300000 400900 50O0OO SOOO09 700000 800000 900000 1000000

Total Number of Nodes

Figure 5-8: ^-^ ploted versus of total number nodes

75

Coarse Intermediate Fine

<$ A «*• <\ <,?> fr> nj) <& A% <£ <?• <y N* a r T v* «r V v

o- a- o- o- <s- is- o- cr o- s-

Mach Number

Figure 5-9: Mach contours for the three grid densities

76

the resulting total pressure losses are plotted in Fig. 5-8. It is apparent that grid dependence

is obtained with a difference in total pressure losses of 1.4% between the intermediate and

fine grids. Figure 5-9 further demonstrates that grid dependence is obtained by viewing the

Mach contours on the symmetry plane of the three grids. Examination of Fig. 5-9 shows

that the coarse grid fails to pick up the flow separation in the exhaust duct and stub along

the inner bend which is present in the other grids. Figure 5-9 also demonstrates that a lower

magnitude of flow separation along the exhaust duct center body is obtained in the coarse

grid where the intermediate and fine grids both demonstrate that flow is separating at the

same order of magnitude. Based on these results the remainding grids in the current study

have been created using a maximum tetrahedral size of h/Q with a computation savings of

nearly 70% less nodes than what was used in the fine grid.

77

Chapter 6

CFD-Based Parametric Study

The following sections present the findings of CFD analysis. Each geometric and aerody­

namic parameter is presented and it is demonstrated, qualitatively and quantitative, how

they are related to the total pressure loss in the exhaust duct. Cross-sections are defined

throughout the study according to the method discussed in Chapter 4 to evaluate and

compare the losses between each duct studied. The sections presented in Fig. 6-1 will be

referenced throughout this Chapter, and distance is calculated along the gas path spline.

13 (Exit)

12 11 10 9 8 7 6 \ / / I

5 \ / /

3 2 1 (Inlet)

Figure 6-1: Cross-sections used for data reduction and flow visualization

Path Spline

78

Section # 1 (Inlet) 2 .3 4 5 6 7 8 9 10 11 12 13 (Exit)

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25

Distance (in)

Figure 6-2: Total pressure loss coefficient evaluated at each section defined in figure 6-1

6.1 Effect of Swirl

Inlet swirl has been demonstrated to have a large influence on the internal flow structure

producing higher losses as inlet swirl is increased. As indicated in Fig. 6-2, the losses are

at a minimum for nominal 0° inlet swirl for a sample exhaust duct with a C-shaped stub.

In this figure, the total pressure loss coefficient is calculated at each section from Fig. 6-1

demonstrating how the total pressure losses can be seen to accumulate. To illustrate this,

contours of the normal component of velocity with vectors of the tangential component is

plotted on cross-sections for the sample exhaust duct in Figs. 6-3 and 6-4 for inlet boundary

conditions of nominal 0° and 35° inlet swirl with low inlet blockage.

The increasing loss between the inlet and section 2 can be attributed to the higher

velocity gradients. A symmetric flow field can be seen at section 2 for inlet swirl of nominal

0° where when inlet swirl is higher as in the case of nominal 35° the flow field is asymmetric.

This asymmetry in flow is caused as a result of an upstream influence of the flow splitter.

The resulting flow field has larger velocity gradients at higher inlet swirl conditions where

79

Normal Velocity [ftsA-l]

Inlet

Figure 6-3: Velocity contours normal to cross-section (nominal 0° swirl)

80

Figure 6-3: Velocity contours normal to cross-section (nominal 0° swirl)....con't

81

<£ <& v <P <ofo v c? <6fo v <$> «° v <r & v f t ? <v «° V> 1? nj> y t? tP Y & & y <£ <& <£> «P

Normal Velocity [ftsM]

Inlet

Figure 6-4: Velocity contours normal to cross-section (nominal 35° swirl)

82

Figure 6-4: Velocity contours normal to cross-section (nominal 35° swirl)....con't

83

viscous shear produces increasing pressure losses.

A large increase in losses is always found to occur between sections 2 and 3. It is at

these sections that the exhaust duct begins transitioning from annular to rectangular and

is where gradients in velocity are highest. Observing Figs. 6-3 and 6-4 at section 3 shows

that inlet swirl transforms the flow field and introduce regions of peak velocity and flow

separation due to incidence angles with the flow splitter. The losses rise at section 3 with

increasing swirl due to increasing asymmetry in the flow field and increasing flow separation.

Sections 3 through 12 finishes the exhaust duct transition from annular to rectangular,

and ends at section 12 bringing the flow out in a direction 90° from the inlet. The total

pressure losses through these sections are found to increase steadily for all inlet swirl con­

ditions as indicated in Fig. 6-2. In Fig. 6-3 for an inlet swirl of nominal 0° high velocity

flow remains concentrated in the lower portion of section 3 near the hub and then separates

off the centerboby at section 5 in a manner similar to Von Karman Vortex shedding. This

low velocity flow can be seen to penetrate in to the core flow field, and is dissipated by

section 12. Counter rotating vortices are seen to form in the core flow by section 8 due to

the fluid turning while the low momentum fluid in the boundary layers at the outer sides of

the section is seen moving opposite to the vortices due to low pressure at the inner surface

of the bend. For the same sections with inlet swirl as in Fig. 6-4, the flow field is more

complex. At section 3 the flow is already seen to be separating at the hub and is joined

at section 4 with the separated flow off the flow splitter. High momentum flow is seen

concentrated along the left hand side of sections 3 through 9 until it becomes fairly more

uniform by section 10. The low velocity pocket of separated flow observed at section 4 is

seen to leave the centerbody at section 5 and can be seen to be almost fully dissipated by

section 12. The effect of turning the flow 90° is again shown by the presence of two counter

rotating vortices which are almost fully developed by section 9.

The remaining duct losses are occurring through the exhaust stub between section 12

through 13 where in both Figs. 6-3 and 6-4 display flow separation along the inner surface

due to the momentum of the fluid being pressed along the outer surface in the remaining

90° turn. The addition of inlet swirl produces a exhaust stub exit flow structure slightly

asymmetric to what is observed with nominal 0° inlet swirl.

84

Straight S-Shape

j $

^ i> <) t? ^ (J (? fl> (J d> «> <J eP «J iJ (? « ° o? Ob V V \ * "? T? 1,* f - ? V N1 (? If «?

Q- Q- Q- <&' ^- O" Q- ©• O- O' O' Cl (V O* Q- <V

Mach Number

Figure 6-5: Mach contours at mid plane demonstrating the effect of stub direction (nominal 0° inlet swirl)

6.2 Effect of Stub Direction

The exhaust stub direction has been shown to have an impact on the internal flow structure

in the exhaust duct. Figure 6-5 demonstrates through Mach contours the effect of stub

direction at nominal 0° inlet swirl on the internal flow structure. At the nominal inlet swirl

of 0° there is more flow separation in the case of the Straight stub and the S-shaped stub

then what is found in the C-shaped stub. The reason for this is seen in Fig. 6-6 where static

pressure contours are presented. In the case of the C-shaped stub, a high pressure region

is present along the outer curvature of the duct which prevents flow separation along the

inner curvature by forcing the flow to follow the surface. As for the case of the Straight and

S-shape stubs, this high pressure region is not present to prevent flow separation as seen in

Fig. 6-6.

85

C-Shape Straight S-Shape

<$ <\N &> \* «J> A & <§> AN t ? \* <& <^ >S> <S° $> 0 s s* V IT V & <r V <b V A* "S7 «T Sr

N<V NQ- 4>' N<V NQ' „$>' ,JV SQ- N 0 ' , $ ' .JV N 0 ' ^" .{V „ y

i««Hi

Q inlet

Figure 6-6: Normalized static pressure contours at mid plane demonstrating the effect of stub direction (nominal 0° inlet swirl)

86

1 (Inlet)

0.45 •

Section # 2 3 4 5 6 7 8 9 10 11 12 13 (Exit)

0.35

Stub Shape

( • ) C-Shape

( • ) S-Stiape

( A ) Straight

I

Distance (in)

Figure 6-7: Total pressure losses calculated for C-shape, Straight, and S-shaped exhaust stubs (low inlet blockage)

The exhaust duct total pressure losses at the exhaust stub exit plane are dependent on

the stub direction. However, the total pressure losses from exhaust duct inlet to the stub

inlet are nearly independent of the stub direction. The graph in Fig. 6-7 demonstrates the

general trend found in this study for a sample exhaust duct. For an inlet swirl angle of 0°

there is no discernable difference in the total pressure loss coefficient calculated from the

inlet up to the stub inlet at section 12. After the stub inlet, there is a clear difference in the

duct losses where the C-shaped stub demonstrates the lower losses with the next highest

being a Straight stub followed by the S-shaped stub being the worst. The same trend is

seen for higher inlet swirl conditions where variations in total pressure losses at a given

section can be less then 5% up to the stub inlet.

6.3 Effect of ECDA

As demonstrated in Figs. 6-3 and 6-4, the flow structure in the annular to rectangular

transition region can be quite complex. A two dimensional flow field becomes asymmetric

87

Mach Number

Figure 6-8: Mach number contours at mid-plane demonstrating the effect of ECDA (inlet conditions: nominal 0° swirl with low blockage)

by the exit of the annular region at section 2 with flow separation occurring in the down

stream sections. To produce lower losses downstream of the annulus, it is desirable to reduce

the fluid Mach number as much as possible.

The effect of the ECDA tested on the flow field can be seen in Fig. 6-8, where Mach

contours are plotted at mid plane of three exhaust ducts containing the different values of

ECDA tested. It is apparent in the figure that an ECDA of 10° led to the lowest mach

numbers at the exhaust duct exit, with the highest occurring when no diffusion is present.

The effect of too much diffusion in the annulus is seen in the figure, where flow separation

is observed along the outer annulus walls of the duct with an ECDA of 20°.

The total pressure losses are calculated at the exit of the annulus of the ducts tested

under inlet swirl conditions with low inlet blockage and presented in Fig. 6-9. The minimum

losses are found to occur in all ducts with an ECDA of 10°. A trend of pressure losses

increasing with increasing inlet swirl is also present in the figure with the exception of ducts

G and H which both have ECDA of 20°. For these two ducts flow the flow separation is

88

0 S 10 15 20 25 30 35 40

Inlet Swirl f°)

Figure 6-9: Total pressure losses as a function of ECDA calculated at exit of the annulus

seen to decrease from nominal 25° inlet swirl to 35° inlet swirl as shown by the contours of

wall shear stress in Figs. 6-10 and 6-11.

6.4 Effect of Wedge Angle

As it was explained in Sec. 6.1 the largest increase in losses is always found to occur between

section 2 and 3. It is between these two sections that we find the flow splitter. Figure 6-12

demonstrates the effect of increasing wedge angle on the pressure losses for three exhaust

ducts with the same ECDA upstream of section 2. At nominal 0° inlet swirl all three duct

produce similar losses however as inlet swirl is increased the exhaust duct with an 80° wedge

angle produces lower losses at 25° inlet swirl. To examine why the three exhaust duct are

not similar at an inlet swirl of 25°, Mach contours are presented in Fig. 6-13 at a given

cross-section cutting through the flow splitter. It is evident in the figure that the flow has

separated along the suction surface of the exhaust ducts with wedge angles of 10° and 45°

while the flow remains attached in the duct with the 80° wedge angle.

89

/ ? f f # /• / .* / .# ? <S> cS> # <? <f «F ^ f <*• «P *

<S' V N * N - •%•' " V " ? > ' ' V t . ' V V

' ,J l _J I Wal l Shear

Ipsa

Figure 6-10: Wall shear stress contours (nominal 25° swirl)

90

# cf>* ^ ^ <?* # ^ *?* ^ <$? ^

•3° o e* <r 0s *r <s> <r «* *?* <av

<v y \ - N- y y y v *• v y

Wall Shear Ipsi]

Figure 6-11: Wall shear stress contours (nominal 35° swirl)

91

ft* I

I

0 5 10 15 20 25 30 35 40

Inlet Swirl (°)

Figure 6-12: Total pressure losses as a function of wedge angle calculated at section 3

6.5 Effect of Aspect Ratio and Area Ratio in the Annular to

Rectangular Transition Region

The effect of aspect ratio and area ratio in the annular to rectangular transition region on

exhaust duct performance is difficult to quantify from the results of this analysis, because

they are under the influence of the geometric parameters that precede them (such as ECDA

and Wedge Angle). It can be seen from the test matrix in Table 3.5 that the influence

of the aspect ratio and area ratio can not be demonstrated independently from the other

geometric parameters. Some conclusions can be made when exhaust ducts are compared

under no inlet swirl and low inlet blockage where those preceding geometric influences are

at a minimum. Figures 6-14 and 6-15 present normalized static pressure and Mach number

of four exhaust ducts which represent the combinations of aspect ratio and area ratio which

are at the corners of the envelop tested. The normalized wall static pressure contours in

Fig. 6-14 demonstrate that gradients on the wall are largest when the aspect ratio and area

ratio are at the minimum. This is unfavorable since large pressure gradient will tend to

promote flow separation off the duct surface. The Mach number contour plots at mid plane

92

U.I4

0.12

0.08

Wedge Angle = 10c - >IRO Wedge Angle = 45

Wedge Angle = 80c

cP 4?i ^ of> ^ <? i ° -l1' 1^ «? -J?1 «J A (S i j (!) (? S> \ t t \» T,N f Y V tJ>" t?> V «?> «* <? V1 4?

<v «s- cr <v c»- Q- o- o cr <v o- Q ©• <v o- o '

Mach Number

Figure 6-13: Mach number contour plot on a plane through the flow splitter

93

AR = 1.000 AS = 0.920

AR= 1.350 AS = 0.920

AR = 1.350 AS = 1.380

AR= 1.000 AS = 1.380

# <f & # & ^ ^ <? ^ & & # # ^ #

•H9H

Q inlet

Figure 6-14: Normalized wall static pressure contours demonstrating the effect of aspect ratio and area ratio in the annular to rectangular transition region

94

AR= 1.000 AS = 0.920

AR = 1.350 AS = 1.380

D AR= 1.350 AS = 0.920

l ) AR= 1.000 AS = 1.380

o < £ " # <0» *<"> N** N> <VS C?5 & <£ <& \* t £ A^ C\Q

ti5 ^ <y <s° v N* *o -^ v" -v* T> v "r v t? <b' <y o- cr <a* <v <y O* Q* O O- <y <S- <y O-

I L.

Mach Number

Figure 6-15: Mach number contours demonstrating the effect of aspect ratio and area ratio in the annular to rectangular transition region

95

in Fig. 6-15 support this theory demonstrating that small aspect ratios do lead to flow

separation off the inner surface of the exhaust duct reducing the available flow area in the

streamwise direction. It can be concluded that it is favorable to have the largest available

aspect ratio to maintain an attached flow.

6.6 Effect of Inlet Boundary Layer Blockage

It has been found that the total pressure losses increase as a result of increasing inlet

boundary layer blockage. As seen in Fig. 6-16 a thick inlet boundary layer is more prone to

flow separation with the point of separation progressing closer to the inlet for thicker inlet

boundary layers. These results where expected when flowing against a positive pressure

gradient. The exhaust ducts with 20° ECDA demonstrated to have the most severe flow

separation over the other ECDA tested. The global effect of inlet blockage on all exhaust

ducts tested can be fully seen the Figs. 6-17 and 6-18. With low inlet blockage, there

no distinct relation to other tested geometric parameters, however increased blockage levels

have demonstrated higher levels of losses related to large inlet diffusion angles. It is apparent

in both figures that for medium and large inlet blockage, the exhaust ducts tested with 20°

ECDA have an overall higher trend of total pressure losses. When comparing medium to

high inlet blockage the same loss trends are found only at higher levels.

96

Inlet blockage

Low(3<1%)

I Medium (4% < B < 8%) High (8 2:10%)

ECDA = 0°

ECDA = 20°

en m (3

4l ECDA =10°

a ri Cm

C$ 1 ^ <?> 0$> «> t?> AN C? I 0 * «? o£> • > «?> V s C$ (? (i> ^ Q* V N5, <> f 'V' f f> V T V >°

c>- <v &• o- <a- o- Q- o- O' o- <y c>- cr o- <y J I ' ' '

Mach Number

Figure 6-16: Mach number contours demonstrating the effect of inlet blockage on three exhaust ducts with ECDA of 0°, 10°, and 20°

97

^

K vs. Inlet Swirl Low and Medium Inlet blockage (p)

0.400 :

0.350'

0.300'

0.260 • I

0.200 I

0.150 i I

0.100

1

0.050 •

0.000 :

• 1

1

0 > a

1 1

• •

0

5 t

a A

ECDA ECDA ECDA

= n° =10° =20°

B<1° o

D

• ! o

* • i D

A

4°<B<8° • 4 •

20 30 35 40

Inlet Swirl f)

jure 6-17: Total pressure loss coefficient vs. inlet swirl for low and medium inlet blockage

OH

I

3!

a,

J » 0.250

II

0.200 11

0.150

0.100

0.050

0.000

K vs. Inlet Swirl Medium and High Inlet blockage (p)

• I

• D

I

1

•! • " • " - " ' 1 ' ' | 1

• D

1 * o

ECDA ECDA ECDA

_

= 0° =10° =20°

,

£

zz

$m: 0

a D

• • D

i

£ A

B>10° • * •

Inlet Swirl (°)

jure 6-18: Total pressure loss coefficient vs. inlet swirl for medium and high inlet blockage

Chapter 7

Correlation of the Total to Total

Pressure Loss

A correlation is presented in the following sections where the total to total pressure loss

coefficient is related to the geometric and aerodynamic parameters based on the data pro­

duced in the computational flow simulations. The resulting correlation serves as a suitable

tool for designers when involved in the preliminary design of annular to rectangular exhaust

ducts.

The results of the parametric study demonstrate that the geometry of the exhaust duct

produces a highly asymmetric flow field, under the influence of swirl, and flow separation is

almost unavoidable when the later is combined with inlet blockage. The favorable approach

to produce the loss correlation would be to build upon an exact solution of the Navier Stokes

equation which can be easily be interpreted by a designer. However, an exact solution to

the Navier Stokes equations does not exist for this complex geometry therefore a numerical

approach must be taken.

The development of a loss model can follow one of the following two approaches; 1)

a purely mathematical approach where a reduced order model or a surrogate model is

constructed from available data (eg. Artificial Neural Networks [25]), and 2) a physical

approach where a numerical correlation is composed of different terms each of which rep­

resent the effect of one flow feature contributing to the loss. The first approach is a black

99

11 vs. AR only those points whose B1 is less than or equal to 0.06

• T1_

• I s If

#>

• '

t f •

eta = 0.72 + 3e*(-0.9 - 1.SAR)

1 1 1 1 0 1 2 3 4 5 6 7 8

AR

Figure 7-1: Diffuser effectiveness versus area ratio with low level inlet aerodynamic blockage and no inlet swirl [7]

box to the designer and does not provide the user with the same feedback as do physical

correlations produced by curve fitting. Through curve fitting, the correlation will provide

the user with a visual understanding of the functional relations which can even take on a

physical meaning of the data being analyzed. The approach taken in this work is the second

one, where the data is correlated through a curve fitting technique similar to what was used

by Japikse [7] where he correlated annular diffuser effectiveness using available published

data.

7.1 Japikse Correlation of Annular Diffusers

To start the process Japikse has first identified that area ratio is the dominant variable

related to diffuser effectiveness, once the data was screened for blockage, and noting that

inlet swirl has been to some extent taken care of through the definition of Cpi in equation

2.3. He then plotted the data versus area ratio, Fig. 7-1, and discovered that the data

followed an exponential trend.

With the equation found in Fig. 7-1, Japikse was able to move on to the other variables

100

r\ vs.«,

a. <

>

« 1 1

1

• b -

• • • • •

" ^ " — • - • J '

I I I

eta= 1.O5-O.M0W1.9

\t * <t

eta = 1.05-0. «)02«,*Z1

'•

i »

<>

25

«1

Figure 7-2: Diffuser effectiveness with principle geometric effects removed including data at all levels of inlet aerodynamic blockage [7]

by removing the effect of area ratio by dividing out the data by the new expression as

presented in Fig. 7-2. In this figure, the data has been plotted versus inlet swirl and has

revealed two trends. The upper trend represents diffusers experiencing mild stall while the

lower trend represents diffusers with substantial stall.

The data was again divided by the new equations defined in Fig. 7-2, and the with the

effects of area ratio and inlet swirl removed from the data Japikse then moved on to inlet

blockage as shown in Fig. 7-3. From this figure Japikse has determined that there are two

trends which have been defined. The lower trend (common blockage, "A"), which is seen

to passes through the square symbols, is data from Coladipietro et al. [18] where tests were

conducted at two different blockage levels. In these testes, inlet conditions consisted of a

clean uniform velocity profile where only the boundary layer thickness was varied. The upper

trend (classical profile blockage, "B") which shows that diffuser performance improves with

increasing inlet blockage comes from inlet conditions where the boundary layer becomes

fully developed and contains increase levels of turbulence and vorticity. Dividing the data

again by the new equations, Japikse represented the data versus inlet blockage in Fig. 7-3

101

il vs. Bi

Jt EC A or

ilJO

0.2

• ^ j

• ^ " • N I

ata- ' i . iK+ o.UB'in(b5l)

o

it

|y = 4.777364E+D1x i -1.217600E+01K +139214BE+Do|

r— „ B »

* " "A"

<>

o.oe

B1

Figure 7-3: Diffuser effectiveness with the principle effects of geometry and inlet swirl removed according to preceding correlations [7]

to reveal that the data has collapsed to a value of 1 ± .10.

The resulting set of equations produced by studying the data trends is the following:

Cp = CPi (ai , r 2 / r i , 62/61) V (AR) 77 ( « I ) r? (£1) (7.1)

rj {AR) = 0.72 + 3e(-o.9-i.5Jifl) ( 7 2 )

r] (ai) = 1.1 - O-OOOlaJ-9 delayed stall (7.3)

r)(ai) = 1.1 -0 .0002a? 1 ear/y staM (7.4)

77 (Bi) = 47.77364B? - 12.17600.Bi + 1.392146 curve 4 , common blockage (7.5)

77 (Si) = 1.22 + 0.08 x ln(Bi) curve B, classical profile blockage (7.6)

Now that the diffuser effectiveness can be predicted, Japikse suggest that a reasonable

first order estimate of the diffuser total pressure loss coefficient can be calculated using Eq.

2.6 (Cp is evaluated from Eq. 7.1 and Cpi from Eq. 2.3).

102

11 vs. Bt

1.2,

1.1

m

•Sf 0.9

"3 •c o.

0.7

i t • *

< •

<

p

< <

"

<

<

>

> ! *

<

r

0.02 0.04 O.OB 0.1 0.12 0.14

Figure 7-4: Diffuser effectiveness with the principle effects of geometry, inlet swirl, and inlet blockage removed [7]

103

K vs. Inlet Swirl Low Inlet blockage (P < 1%)

C-l

=8 Co

• • v .

a, i 1

i

1

4

0.450-

0.400 -

0.350'

0.300

0.250 •

0.200;

0.150

0.100

0.050 •

0.000 :

0 5 10 15 20 25 30 35 40

Inlet Swirl (°)

Figure 7-5: Total pressure loss trend in the exhaust for low inlet blockage

7.2 Correlation of the CFD Results

It was demonstrated in Sec. 6.2 that the total pressure losses in the exhaust duct are

independent of the stub direction. For this reason it was chosen to correlate the exhaust

duct losses independently of the stub losses. The following section presents a corelation to

predict the total pressure losses between the inlet and section-12 as shown in Fig. 6-1. The

correlations produced in this work have been found with the help of Windows Excel and

the statistical package LAB Fit [26].

The fist parameter that demonstrated to have the first order impact on losses in the

exhaust duct is inlet swirl. When inlet blockage was varied some distinct trends appeared

which demonstrated different behaviors in the losses depending on the magnitude of diffusion

taking place in the annular inlet of the exhaust duct. For low inlet blockage (j3 ^ 1%) it

is possible to find a single trend which can be used to normalize the data, Fig. 7-5. For

medium inlet blockage (4% < (3 ^ 8%) one trend has been defined for an ECDA of 20° and

one for the lower values tested, Fig. 7-6. When high inlet blockage (/3 > 10%) is present a

104

0.450

0.400 •

r v j

t '6

•O ft, |

*5 3

**** ft.

3 •5 53,

0.350

0.300

0.250

0.200

* 0.150

0.100

0.050

0.D00

K vs. Inlet Swirl Medium Inlet blockage (4% < p < 3%)

Y »O.112800C«Sff (0.047726-Jf) ECDA 0'

10' 201

20 25 30 35 40

Inlet Swirl f)

Figure 7-6: Total pressure loss trend in the exhaust for medium inlet blockage

8 ft, I

ft,

•5

II

0.400 •

0.350

0.300

0.250

0.200

0.150

0.100

0.050

0.000

K vs. Inlet Swirl High Inlet blockage (p > 10%)

Y - 0.214385COS/{0.033095-A'Xl+#)

J = 0.1 limCOSH (0.047726-XXI+ #)

5=0.1

ECDA • 0° * 10° • 20°

Met Swirl (°)

Figure 7-7: Total pressure loss trend in the exhaust for large inlet blockage

105

Figure 7-8: Surface passing through normalized losses for ECDA and aspect ratio (Solid circles are points lying above the surface and hollow circles are points lying below)

correction factor is used to increase the level of the trends used to describe medium inlet

blockage, Fig. 7-7. To calculate the loss due to intermediate ECDA and blockage levels, it is

recommended to interpolate between the trends in Fig. 7-6 and then interpolated with Fig.

7-5 for smaller inlet blockage or select a smaller correction factor for higher inlet blockages.

The equations derived above have considered the contribution to the losses due to the

effect of swirl with some consideration to the annular inlet section defined by ECDA. After

normalizing the data with those equations the effect of swirl with some consideration to

ECDA have effectively been removed and the other contributing parameters can now be

evaluated.

The next parameter considered is the aspect ratio of the exhaust duct. A good fit to

the data could not be found after trials at correlating the aspect ratio to the normalized

106

1.400

*4

1.200 H

1.000

4q DOT­

S' 8 ft) S 0.600 •

* | 0.400 \

0.200 \

K vs. Wedge Angle

' * QJ0GO3JT1 -• 0.00338* +1.07369

- I

0.000 4 1 -i 1 1 1 1 1 1 1 1 1 1 1 1 r

0 5 10 15 20 25 30 35 40 45 50 55 BO 65 70 75 80 85 90

Wedge Angle {')

Figure 7-9: Normalized losses versus wedge angle with the effects of blockage, swirl, ECDA, and aspect ratio removed

data demonstrating that some other parameter is influencing the losses. It was discovered

that there may still remain some influence of the annular inlet effecting the data and when

correlating ECDA with the aspect ratio a surface could be passed through the data with

satisfactory results, Fig. 7-8. The surface plot in Fig. 7-8 supports the results presented

in Sec. 6.3 and 6.5 demonstrating that it is favorable to have a large aspect ratio and that

there is an optimum ECDA between 0° and 20°.

The data are again normalized with respect to the surface equation defined above and

plotted against the values of wedge angles tested, Fig. 7-9. A trend is visible and a

polynomial fit demonstrates that there is an optimum wedge angle near 60°. Once the data

are normalized again for this trend we see in Fig. 7-10 that there is only minimum influence

of the area ratio on the normalized pressure losses and a polynomial fit is made to the data.

The final result of the correlation can be seen in Fig. 7-11. It is demonstrated that 70%

of the data falls within ±10% error and 85% of the data falls within ±15% error. A more

detailed breakdown of the accuracy of the correlation is shown in Table 7.1. the resulting

107

K vs. Area Ratio

•§ 1.000 S

3 0.600 •

i i t

\r~-

• »

1 * •

= 1 84000-**-O.S330CMT3|

1

k

0.95 1.00 1.05 1.10 1.15 1.20 1.25 1.30 1.35 1.40

Ansa Ratio

Figure 7-10: Normalized losses versus area ratio with the effects of blockage, swirl, ECDA, aspect ratio, and wedge angle removed

Table 7.1: Break Down of Correlation Accuracy

Error (\(KCFD - KCorreiation\) Data Points(%) 0.00-0.02 0.02-0.04 0.04-0.06

72 22 6

108

CFD vs. Predicted

0.00 0.10 0.20 0.30 0.40 0.50

J V [Swtrl[[8CDA,AspectRatio][Wedge][AreaRatio]

Figure 7-11: Plot of CFD losses vs. predicted losses between inlet and section-12

109

set of equations are as follows:

K = K{Swirl, X blockage)K(ECDA, AspectRatio)K(WedgeAngle)K(AreaRatio) (7.7)

K(Swirl, low blockage) = 0.101720 cosh(0.047854a) all ECDA (7.8)

K(Swirl, medium blockage) = 0.112800 cosh(0.047726a) up to 10° ECDA (7.9)

K(Swirl, medium blockage) = 0.214385 cosh(0.033095a) 20° ECDA (7.10)

K(Swirl, high blockage) = K(Swirl, medium blockage)(l + £) £ = 0.1 (7.11)

( 0 78730 \

- — — 0.07885ECDA ) +0.0398OECZM AspectRatio )

(7.12)

K (WedgeAngle) = 0.00003(WedgeAngle)2 - 0.00338(Wedge Angle) + 1.07369 (7.13) K(AreaRatio) = 1.84000(Ar eaRatio) - 0.8Z300(AreaRatio)2 (7.14)

110

Chapter 8

Conclusion and Recommendation

8.1 Conclusion

A parametric CFD study has been carried out where key geometric and aerodynamic pa­

rameters are varied and the effect on the total pressure loss is observed. The results of the

CFD data were then used to produce a correlation to predict the duct losses. Based on this

work the following conclusions are made:

1. Inlet Swirl

• Inlet swirl was demonstrated to have a high order effect on the pressure loss

in the exhaust duct. The minimum pressure losses occur when the swirl is at

a minimum. The introduction of inlet swirl causes an asymmetric flow field

yielding to flow separation with increased pressure losses.

2. Stub direction

• pressure loss at the exhaust duct exit was found to be independent of the stub

direction. The highest losses occur in the S-shaped stubs as a results unfavorable

flow separation due to pressure gradients.

3. ECDA

• The degree of ECDA demonstrated that too little and too large of an ECDA is

not favorable. The optimum value of ECDA resulting in the minimum velocity at

111

the inlet to the annular to rectangular transition region without flow separation

is between 7° and 10°.

4. Wedge Angle

• The three wedge angles tested perform similarly for inlet swirls up to 25°. At

this swirl angle flow separation on the suction surface of the flow splitter occurs

on the wedge angles of 10° and 45° and is delayed for the 80° flow splitter. At

35° inlet swirl the 80° wedge flow splitter would separate on the suction surface.

It is desirable to have a large wedge angle for reduced pressure loss.

5. Aspect Ratio and Area Ratio

• It was difficult to quantify the effect of these parameters since they proceed the

other geometric parameters. A small aspect ratio and area ratio lead to pressure

gradients that can result in flow separation.

6. Inlet Boundary Layer Blockage

• The total pressure losses increase with increasing inlet blockage. Exhaust ducts

with ECDA of 20° performed to worst of the geometries tested for medium and

high inlet blockage.

7. Correlation of the CFD Results

• The Taguchi Design provided a test matrix that provided the minimum amount

of exhaust duct models to successfully capture the resulting effects on the total

pressure loss coefficient. The CFD data has been correlated to the total pressure

loss coefficient with reasonable accuracy.

8.2 Recommendation

Examining the results of the current study, the following recommendations are made for

the loss modeling of non-symmetric gas turbine exhaust ducts using CFD:

112

1. As future work, it is desirable to include the exhaust stubs in the correlation. To

do this, first, a survey should be conducted to define common exhaust stubs used

in turbo-machinery applications which contain non-symmetric gas turbine exhaust

ducts. Secondly, model them numerically as in the current work and correlate the

pressure losses.

2. The correlation produced from the current work serves as a guide for designers to pro­

duce more efficient non-symmetric gas turbine exhaust ducts. To be able to accurate

predict the total pressure loss will require that the current CFD work be calibrated

to experimental data.

113

Bibliography

[1] G. Sovran and E. Klomp, "Experimentally determined optimum geometries for recti­

linear diffusers with rectangular, conical , or annular cross-section," Fluid Mechanics

of Internal Flows, pp. 270-319, 1967.

[2] A. McDonald and R. Fox, "An experimental investigation of incompressible flow in

conical diffusers," ASME Paper, no. 65-FE-25, 1965.

[3] J. H. G. Howard, A. B. Thornton-Trump, and H. J. Henseler, "Performance and flow

regimes for annular diffusers," ASME Paper, no. 67-WA/FE-21, 1967.

[4] D. Japikse and N. C. Baines, Diffuser Design Technology. Concepts ETI, Inc., 1998.

[5] A. Elkersh, A. Elgammal, and N. Maccallum, "An experimental investigation of the

performance of equiangular annular diffusers with swirled flow," Proceedings of the

Institution of Mechanical Engineers, vol. 199, no. C4, pp. 293-297, 1985.

[6] A. Klein, "Characteristics of combustor diffusers," Prog Aerospace Sci., vol. 31,

pp. 171-271, 1995.

[7] D. Japikse, "Correlation of annular diffuser performance with geometry, swirl, and

blockage," Proceedings of the 11th Thermal and Fluids Analysis Workshop, pp. 107-

118, 2000.

[8] M. Cunningham, Flow in Non-Symmetric Gas Turbine Exhaust Ducts. PhD thesis,

Queen's University, 2002.

114

[9] A. MacDonald, R. Fox, and R. Van Dewoestine, "Effects of swirling inlet flow on

pressure recovery in conical diffusers," AIAA Journal, vol. 9, no. 10, pp. 2014-2018,

1971.

[10] Y. Senoo, N. Kawaguchi, and T. Nagata, "Swirl flow in conical diffusers," Bulletin of

the JSME, vol. 21, no. 151, pp. 112-119, 1978.

[11] F. Dolan and P. J. Runstadler, "Pressure recovery performance of conical diffusers at

high subsonic mach numbers," NASA CR-2299, 1973.

[12] V. K. Sharan, "An exponential investigation of the behavior of conical diffusers in

turbulent flow," Journ Appl Math and Phys, vol. 27, 1976.

[13] J. Livesey and A. Odukwe, "Some effects of pipe flow generated entry conditions on

the performance of straight walled conical diffusers with high subsonic entry mach

number," Israel Journal of Technology, vol. 11, no. 4, pp. 217-222, 1973.

[14] D. Japikse, "A new diffuser mapping technique, part 1: Studies in component perfor­

mance," Transactions of the ASME, vol. 108, pp. 148-156, 1986.

[15] A. Klein, "Review: Effects of inlet conditions on conical diffuser performance," ASME,

Transactions, Journal of Fluids Engineering, vol. 103, pp. 250-257, 1981.

[16] S. Dovzhik and V. Kartavenko, "Measurement of the effect of flow swirl on the efficiency

of annular ducts and exhaust nozzles of axial turbomachines," Fluid Mechanics - Soviet

Research, vol. 4, no. 4, pp. 156-172, 1975.

[17] E. Klomp, "Performance of straight-walled annular diffusers with swirling flow," Aero­

nautical Journal, vol. 101, no. 1010, pp. 467-480, 1997.

[18] R. Coladipietro, J. Schneider, and K. Sridhar, "Effects of inlet flow conditions on the

performance of equiangular annular diffusers," Transactions of the CSME, vol. 3, no. 2,

pp. 75-82, 1975.

[19] S. Stevens and G. Williams, "The influence of inlet conditions on the performance

of annular diffusers," ASME, Transactions, Journal of Fluids Engineering, vol. 102,

pp. 357-363, 1980.

115

[20] M. Loka, M. Robichaud, W. Di Bartolomeo, D. Loe, and D. Sowers, "Aerodynamic op­

timization of the exhaust-ejector system on the tiltrotor using navier-stokes analysis,"

paper AIAA 2000-0986, AIAA, 2000.

[21] R. K. Roy, Design of experiments using the Taguchi approach : 16 steps to product and

process improvement. Wiley, 2001.

[22] D. Wyatt, "Analysis of errors introduced by several methods of weighting nonuniform

duct flows," Tech. Rep. TN 3400, National Advisory Committee on Aeronautics, 1955.

[23] C. Rhie and W. Chow, "A numerical study of the turbulent flow past an isolated airfoil

with trailing edge separation," AIAA Paper 82-0998, 1982.

[24] F. Menter, "Zonal two equation k-u turbulence models for aerodynamic flows," AIAA

Paper 93-2906, 1993.

[25] R. J. Schalkoff, Artificial Neural Networks. McGraw-Hill, 1997.

[26] W. P. Silva and C. M. D. P. S. Silva, LAB Fit Curve Fitting Software (Nonlinear

Regression and Treatment of Data Program) V 7.2.42, 1999-2008.

116


Recommended